TRANSGENIC CONTROL OF THE CITRUS WEEVIL,...

126
1 TRANSGENIC CONTROL OF THE CITRUS WEEVIL, DIAPREPES ABBREVIATUS: A THREE-PRONGED APPROACH TARGETING DIGESTION, GUT PHYSIOLOGY, AND GENE EXPRESSION By SULLEY KWEKU BEN-MAHMOUD A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2013

Transcript of TRANSGENIC CONTROL OF THE CITRUS WEEVIL,...

1

TRANSGENIC CONTROL OF THE CITRUS WEEVIL, DIAPREPES ABBREVIATUS: A THREE-PRONGED APPROACH TARGETING DIGESTION, GUT PHYSIOLOGY,

AND GENE EXPRESSION

By

SULLEY KWEKU BEN-MAHMOUD

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2013

2

© 2013 Sulley Kweku Ben-Mahmoud

3

To the Ben-Mahmoud family, especially my wife, Theodora

4

ACKNOWLEDGMENTS

Above all, I would like to thank my dissertation committee Ronald D. Cave,

Robert G. Shatters Jr., Charles A. Powell, Daniel A. Hahn, and James E. Maruniak. I

am most grateful to R. D. Cave for his comments and guidance. For his advice,

encouragement, and intellectual stimulation, I am profoundly grateful to R. G. Shatters

Jr. I thank C. A. Powell for his support. I cannot go on without extending my sincerest

appreciation to Dov Borovsky for being a great teacher and mentor. I feel fortunate to

have had D. G. Hall, S. L. Lapointe, and J. E. Ramos in my camp, particularly for their

expert advice in various aspects of my research.

I am thankful to D. Borovsky, R. G. Shatters Jr., C. A. Powell, D. G. Hall and S. L.

Lapointe for allowing me to build on studies they begun at the University of Florida and

the USDA-USHRL (Fort Pierce). My appreciation goes to H. Van Praet and B. Diego for

their help in organizing the transgenic citrus inventory and to L. Faulkner, J. Smith and

C. Peck for their help in citrus maintenance. The potted-plant bioassay was an arduous

task and I could not have accomplished it without the help of R. G. Shatters Jr., B.

Diego, A. F. Voss, C. A. Malone, S. Clark, R. Jain, and K. Moulton. My special thanks to

A.S. Hill for her help with maintaining the D. abbreviatus colony and to L. Markle for his

help in setting up the larval bioassays. I am also grateful to E. Egan and L. I. Shaffer for

making their help readily available when I needed it. I really appreciate C. L. McKenzie,

W. B. Hunter, G. A. Luzio, R. P. Niedz, J. A. Manthey, R. G. Cameron and M. E. Hilf at

the USDA-USHRL for making their laboratories available to me. I would be remiss if I

forgot to add P. D’Aiuto, S. Kauffman, and E. Branca to my acknowledgments. The

camaraderie and friendship of E. Van Ekert and C. L. Hawkings is appreciated as we

endured the dissertation process.

5

To my friends: Sheri Anderson, Adam Searle, Erik Blosser, Irka Bargielowski,

Rodrigo Diaz, Veronica Manrique, Angie Alejandra Niño, Maud Verstraete, Josh Voss,

Rocco Alessandro, Paul Robbins, Kent Morgan, Pasco Avery, John Prokop, Mary

Prokop and others I have failed to mention: “me da mo nyinara ase!”. I immensely

enjoyed your comradeship.

6

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS .................................................................................................. 4

LIST OF TABLES ............................................................................................................ 8

LIST OF FIGURES .......................................................................................................... 9

LIST OF ABBREVIATIONS ........................................................................................... 11

ABSTRACT ................................................................................................................... 15

CHAPTER

1 INTRODUCTION AND LITERATURE REVIEW ..................................................... 17

Life cycle of Diaprepes abbreviatus in Citrus .......................................................... 17 Current Control of Diaprepes abbreviatus............................................................... 22

Future Control of Diaprepes abbreviatus ................................................................ 24 Transgenic Control of Diaprepes abbreviatus .................................................. 25

Candidate Toxins ............................................................................................. 26 Delivery of Toxins in a Transgenic Strategy ............................................................ 35

2 RECOMBINANT EXPRESSION AND CHARACTERIZATION OF DIAPREPES ABBREVIATUS CATHEPSIN L1 AND THE ROLE OF CATHEPSIN L PROTEASE INHIBITORS IN THE REGULATION OF CATHEPSINS INVOLVED IN DIGESTION. ................................................................................... 37

Materials and Methods............................................................................................ 39 Molecular Modeling and Docking of Cathepsin L Protease Inhibitor to D.

abbreviatus of Cathepsin L1 .......................................................................... 39

Rearing of Diaprepes abbreviatus .................................................................... 41

Bacterial Strains and Plasmids ......................................................................... 41 Primers ............................................................................................................. 41 Cathepsin L 1 and Cathepsin L Protease Inhibitor Constructs ......................... 42

Site-Directed Mutagenesis ............................................................................... 43 Expression of Recombinant Proteins ............................................................... 44 Purification of Recombinant Proteins ............................................................... 46

Enzymatic Assays ............................................................................................ 47 Results .................................................................................................................... 48

Modeling of Cathepsin L1 and CPI1 ................................................................. 48 Synthesis and Purification of DaCatL1 ............................................................. 49 Synthesis and Purification of CPI ..................................................................... 51 Cathepsin L1 Substrate Specificity and pH optimum ........................................ 51 Effect of inhibitors on Cathepsin L1 activity ...................................................... 52

7

Sequence alignment of Cathepsin L1 and Human cathepsins ......................... 53 D. abbreviatus Gut in vitro Assays ................................................................... 53

Discussion .............................................................................................................. 54

3 A METHOD FOR CONTROLLING DIAPREPES ABBREVIATUS LARVAL FEEDING ON CITRUS USING A BACILLUS THURINGIENSIS CYTOLYTIC TOXIN ..................................................................................................................... 69

Materials and Methods............................................................................................ 70 Citrus Plants ..................................................................................................... 70

Primers ............................................................................................................. 71 Nucleic Acid Extractions ................................................................................... 71 Quantitative Polymerase Chain Reaction (qPCR) ............................................ 74

Quantitative Reverse Polymerase Chain Reaction (q-RT-PCR) ....................... 74 Potted-Plant Bioassays .................................................................................... 75

Results .................................................................................................................... 78

Potted-Plant Bioassays ........................................................................................... 79 Discussion .............................................................................................................. 81

4 TARGETING THE PERITROPHIC MEMBRANE OF DIAPREPES ABBREVIATUS VIA RNAi INDUCES A DOSE-DEPENDENT MORTALITY RESPONSE ............................................................................................................ 94

Introduction ............................................................................................................. 94 Materials and Methods............................................................................................ 97

Diaprepes abbreviatus Larvae .......................................................................... 97

Primers ............................................................................................................. 97

Double-stranded RNA Synthesis. ..................................................................... 97 Larval Bioassay ................................................................................................ 99

Results .................................................................................................................. 100

Discussion ............................................................................................................ 101

5 GENERAL CONCLUSIONS AND FUTURE DIRECTIONS .................................. 105

LIST OF REFERENCES ............................................................................................. 112

BIOGRAPHICAL SKETCH .......................................................................................... 126

8

LIST OF TABLES

Table page 3-1 Primers used for qPCR and q-RT-PCR .............................................................. 71

3-2 Summary statistics of Ct values obtained from q-RT-PCR analysis of Cyt2Ca1 citrus and non-engineered citrus ......................................................... 90

3-3 Statistical analysis of uninfested root ratio .......................................................... 91

3-4 Statistical analysis of infested root ratio .............................................................. 91

3-5 Statistical analysis of root damage index ............................................................ 92

3-6 Statistical analysis of weight gain of surviving larvae fed on citrus ..................... 92

3-7 Statistical analysis of larval mortality .................................................................. 93

4-1 Primers for Diaprepes abbreviatus dsRNA synthesis. ........................................ 98

9

LIST OF FIGURES

Figure page 1-1 General plant and root damage as a result of Diaprepes abbreviates feeding

on transgenic versus wild-type alfalfa. ................................................................ 28

1-2 Amino acid sequence alignments of cathepsin L protease inhibitors (CPI 1 and 2) to cathepsin L 1 and 2 in Diaprepes abbreviatus. .................................... 32

2-1 Modelling of the interaction between Diaprepes abbreviatus cathepsin L1 (DaCatL1) and cathepsin L protease inhibitor 1 (CPI1). ................................. 6060

2-2 SDS-PAGE analysis of recombinant cathepsin L1 purification. .......................... 61

2-3 Western blot analysis of His- tagged DaCatL1. .................................................. 62

2-4 Mass spectrometry analysis of DaCatL1. ........................................................... 62

2-5 Size-exclusion purification of DaCatL1. .............................................................. 63

2-6 SDS PAGE (lanes a-d) and western blot analysis (lane e) of cathepsin L protease inhibitor1 mutant (CPI1mut). ................................................................ 63

2-7 Fluorogenic substrate preference of DaCatL1.. .................................................. 64

2-8 pH activity profile of recombinant DaCatL1 using Z-FR-AFC.............................. 64

2-9 Inhibition of DaCatL1 by small-molecule protease inhibitors. ............................. 65

2-10 Inhibition of cathepsin L protease inhibitors on the activity of DaCatL1. . .......... 65

2-11 Alignment of DaCatL1 with human cathepsin sequences. Highlighted in yellow are the active site residues. ..................................................................... 66

2-12 Influence of TritonTM X-100 (1%) on the interaction between D. abbreviatus gut cathepsins and the cathepsin L protease inhibitors CPI1 and CPI1mut. ...... 67

2-13 Inhibition of D. abbreviatus larval cathepsins by cathepsin L protease inhibitors ............................................................................................................. 68

3-1 Experimental design of potted plant bioassay. ................................................... 84

3-2 Relative levels of transcription of Cyt2Ca1 gene in transgenic citrus plants normalized with the 18S RNA levels. .................................................................. 85

3-3 Some transgenic citrus cuttings used for potted plant bioassays. ...................... 86

3-4 Resistance of Cyt2Ca1 engineered citrus versus wild type control. .................. 87

10

3-5 Effect of larval root feeding on Cyt2Ca1 citrus plants and the concomitant effect on larval weight and mortality. .................................................................. 88

4-1 Corrected mortality (post 12 days) due to feeding on dsRNA sequences targeting Diaprepes abbreviatus genes. ........................................................... 104

11

LIST OF ABBREVIATIONS

Act Actin

AFC 7-Amino-4-trifluoromethylcoumarin

AMC 7-amino-4-methylcoumarin

ANOVA Analysis of Variance

Bt Bacillus thuringiensis

CA074 (L-3-trans-(Propylcarbamyl)oxirane-2-carbonyl)-L-isoleucyl-L-proline methyl ester

CAL2 Cathepsin L2

CD Citrus dehydrin

CHAPS 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate

CPI Cathepsin L Protease Inhibitor

CPI1 Cathepsin L Protease Inhibitor 1

CPI1mut Cathepsin L Protease Inhibitor 1 mutant

Ct Cycle Threshold

CTV Citrus Tristeza Virus

DaCatL1 Diaprepes abbreviatus Cathepsin L1

DC Diaphorina citri

DEPC Diethylpyrocarbonate

DNA Deoxyribonucleic Acid

dsRNA double-stranded RNA

DTT Dithiothreitol

E. coli Escherichia coli

E-64 L-trans-Epoxysuccinyl-leucyamido (4-guanidino) butane

EDTA Ethylenediaminetetraacetic acid

FOXO forkhead transcription factor

12

FR Phenylalanine and Arginine

GPR Glycine, Proline, and Arginine

GST Glutathione S-transferase

HE High Expresser

His6 Polyhistidine Tag

HumCTB Human Cathepsin B

HumCTH Human Cathepsin H

HumCTK Human Cathepsin K

HumCTL1 Human Cathepsin L1

HumCTS Human Cathepsin S

IPM Integrated Pest Management

IPTG Isopropyl β-D-1-thiogalactopyranoside

IRR Infested Root Ratio

LE Low Expresser

MBP Maltose Binding Protein

ME Medium Expresser

MEROPS Peptidase Database

MES 2-(N-morpholino)ethanesulfonic acid

mRNA Messenger RNA

MS Mass Spectrometry

MW Molecular Weight

OGNC Octyl Glucose Neopentyl Glycol

PAGE Polyacrylamide Gel Electrophoresis

PCR Polymerase Chain Reaction

PMCBP Peritrophic Membrane Chitin Binding Protein

13

PMP Peritrophic Matrix Protein

pNA p-nitroanilide

PP Propeptide

qPCR Quantitative Polymerase Chain Reaction

q-RT-PCR Quantitative Reverse Transcriptase Polymerase Chain Reaction

RDI Root Damage Inndex

RdRP RNA dependent RNA polymerase

RFU Relative Fluorescent Unit

RISC RNA Interfering Silencing Complex

RNA Ribonucleic Acid

RNAi RNA interference

rRNA Ribosomal RNA

SDS Sodium Dodecyl Sulfate

SDSC San Diego SuperComputer Center

SEC Size Exclusion Chromatography

siRNA Small Interfering RNA

S-S Bond Disulfide Bond

Suc-AAPF N-succinyl-alanine-alanine-proline-phenyalanine-p-nitroanilide

SUMO Small Ubiquitin-like Modifier

TAE Tris, Acetic Acid, EDTA

Tc-ASH Tc-achaete-scute-homolog

TMOF Trypsin Modulating Oostatic Factor

TPCK Tosyl phenylalanyl chloromethyl ketone

Ubi Ubiquitin

UF-ICBR University of Florida-Interdisciplinary Center for Biotechnology Research

14

URR Uninfested Root Ration

USDA-USHRL United States Department of Agriculture-United States Horticultural Research Laboratory

VVR Valine, Valine, and Arginine

WT Wild-Type

Z Benzyloxycarbonyl

Z-FY(tBu)DMK (N-benzyloxycarbonyl-phenylalanyl-t-butyl-tryrosyl diazomethylketone

α-Tub Alpha Tubulin

βME Beta-Mercaptoehtanol

15

Abstract of Dissertation Presented to the Graduate School

of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

TRANSGENIC CONTROL OF THE CITRUS WEEVIL, DIAPREPES ABBREVIATUS: A THREE-PRONGED APPROACH TARGETING DIGESTION, GUT PHYSIOLOGY,

AND GENE EXPRESSION

By

Sulley Kweku Ben-Mahmoud

December 2013

Chair: Ronald D. Cave. Co-chair: Robert G. Shatters, Jr. Major: Entomology and Nematology

Diaprepes abbreviatus (L.) continues to threaten citriculture in the USA. A

transgenic citrus rootstock approach is proposed to manipulate citrus to synthesize

molecules selectively toxic to D. abbreviatus in a pyramiding/stacking strategy to delay

the onset of resistance. Candidate toxins are sought to complement Cyt2Ca1 transgenic

citrus, with cathepsin L protease inhibitors to inhibit protein digestion, and dsRNA to

interfere with gene expression.

The synthesis, purification, and characterization of a cysteine protease,

cathepsin L1 (DaCatL1) from D. abbreviatus are described, in addition to endogenous

inhibitors of cathepsins, cathepsin L protease inhibitor (CPI1) and a mutant (CPI1mut).

Complementary DNA encoding DaCatL1 was cloned into an expression vector fused

with GST and His6 tags for copious expression in an Escherichia coli strain. CPI1mut

was engineered by site-directed mutagenesis of CPI1 by replacing Lys101 of CPI1 with

a Cys residue. DaCatL1 was stable in solution with 1% TritonTM X-100, whereas CPI1

and CPI1mut were synthesized in soluble forms. Enzymatically active DaCatL1 (23KDa)

16

was cleaved from the 60-KDa fusion protein by buffer exchange at alkaline pH using

size exclusion chromatography after initial purification with a Ni-NTA resin. DaCatL1

had optimal activity at pH 8 with the substrate Z-FR-AFC, and enzyme inhibition assays

with protease inhibitors revealed cryptic chemical characteristics. The inhibitors

effectively inhibited cathepsins extracted from the midgut of 6-week-old larvae at acidic

pH but not at alkaline pH. DaCatL1 is not inhibited by either CPI1 or CPI1mut at pH8.

There appear to be two broad classes (acidic and basic) of cathepsins in D.

abbreviatus.

Quantitative PCR of 75 genetically transformed citrus plants identified 31

containing Cyt2Ca1. An ad hoc criterion using q-RT-PCR data was used to select seven

transgenic citrus plants expressing Cyt2Ca1. Three transgenic citrus events expressing

Cyt2Ca1 were significantly more resistant to feeding damage by D. abbreviatus larvae

in no-choice potted-plant bioassays relative to their untransformed wild type cohorts.

Feeding four dsRNA sequences to silence D. abbreviatus genes identified two

possible targets for RNAi. The peritrophic membrane chitin binding protein and

cathepsin L2 dsRNA sequences are possible candidates for a transgenic strategy to

control D. abbreviatus.

17

CHAPTER 1 INTRODUCTION AND LITERATURE REVIEW

Life cycle of Diaprepes abbreviatus in Citrus

Since its accidental introduction into the United States in the mid-1960s, the

citrus weevil, Diaprepes abbreviatus (Linnaeus) (Coleoptera: Curculionidae), has been

a constant threat to the cultivation of citrus. The weevil was first reported in a citrus

grove in Apopka, Florida in 1964 (Beavers and Salhime, 1975). Diaprepes abbreviatus

has since exhibited the ability to spread and increase its pest status. It is currently found

in all the major citrus producing states: Florida, Texas, and California (Lapointe et al.,

2007). It originated in the West Indies (Ernst et al., 2006) where it is an economically

important pest of sugar cane, yams, pineapple, and corn (Hall, 1995). The weevil has a

wide host range, in excess of 300 plant species from 59 plant families that include many

commercially cultivated crops, e.g. corn and potatoes (Simpson et al., 1996). In the

United States, it has established itself as a high-impact, economic pest of citrus. As a

result of its polyphagous feeding habit, D. abbreviatus is a problematic pest to the

cultivation of a number of cash crops and ornamentals (Diaz et al., 2006; Martin et al.,

2009).

Diaprepes abbreviatus has numerous other common names from its association

with a variety of host plants and geographic distribution: Apopka weevil, West Indian

weevil, and sugarcane rootstalk borer weevil are examples. Adults are large and grow

to approximately 2 cm in length, and typically have black elytra with brightly colored

streaks of orange, yellow, red, or gray scales in numerous morphotypes. Adults live

mainly in the crown of trees, where they mate and feed predominantly on leaves

(Beavers and Selhime, 1975). Characteristics of a D. abbreviatus infestation include a

18

notching pattern on the leaf edges and appearance of significant frass on leaves. Upon

emergence as adults from underground (larval and pupal stages are almost entirely

subterranean), they fly for a short distance, seeking a host tree which they hardly leave

unless disturbed or when food is depleted (Weissling et al., 2009). Adults have an

average life-span of 3-5 months, with females usually out-living males.

Females have high fecundity (Sirot and Lapointe, 2008). Mating occurs in the

early hours of the day or at night when the temperature is cool. This is the time during

which they are also most active. Females lay eggs in clusters of 30-250 eggs arranged

as a single layer in no particular pattern. These clusters are laid between two leaves,

which they hold together at the edges with a gelatinous substance that they secrete.

Eggs are oval and approximately 1 mm in length, off-white in color but darken as they

age, and hatch in about a week (7-10 days). Larvae are white and legless, and almost

immediately drop to the ground after hatching and burrow into the soil to reach the roots

of trees. In the soil, they feed voraciously on roots, which is their only known source of

nutrition in the wild.

Neonate larvae commence feeding on fibrous roots and then move to larger

roots as they mature. Larvae complete 10 or 11 instars, with the third to ninth instars (3-

4 months after hatching) the most aggressive feeders. Feeding damage may be severe

enough to kill a citrus tree as a result of girdling; however, feeding injuries serves as an

entry point for soil-borne plant pathogens (Graham et al., 2003). Although larvae feed

copiously on roots, they can also survive for long periods without food. The entire

larval stage may last from a few months to slightly over a year.

19

Pupation occurs in a vertical chamber that the larva (resting with its head up)

compacts by active wriggling movements. Following pupation, there is a two-week

period during which teneral adults remain quiescent, though they may physically appear

to be ready to go above ground. Adults tunnel through soil to the surface by means of a

pair of deciduous mandiblular cusps which often break off. The adults seek mates

almost as soon as they emerge. The developmental period of the citrus root weevil is

highly variable, ranging from six months to as long as two years, depending on nutrition,

temperature, humidity, and soil moisture.

The larva of the weevil is the most damaging stage. It feeds on the roots, is long-

lived, and is almost entirely subterranean. These factors make dependence on

insecticides nearly futile in achieving cost-effective control. Although accurate economic

estimates are hard to come by, annual loss in the citrus industry due to the weevil in

Florida alone in 1997 were estimated to be in excess of $75 million (Stuart et al., 2006).

This is very significant, particularly for an insect with such a wide host range.

No resistant cultivars of citrus are available (Shapiro and Gottwald, 1995). Using

standard plant breeding techniques to breed D. abbreviatus resistance into commercial

citrus varieties has not been viable as yet, though an antixenotic effect has been

observed in two varieties of Poncirus trifoliata L. Raf. in the laboratory (Lapointe and

Bowman, 2002). A need to develop new methods to control the weevil is therefore

relevant. Transgenic citrus strains with specific resistance to D. abbreviatus could play a

key role in protecting citrus. The current literature is replete with numerous examples of

using transgenic techniques to improve crop resistance to pests (Harlander, 2002;

James, 2012); however, almost all transgenic strategies for crop resistance to insects

20

rely on the transgenic expression of a variety of Bacillus thuringiensis (Bt) toxins. The Bt

toxins are pore-forming proteins, with specific insecticidal activities, produced during the

sporulation phase of different strains of the bacteria. Commercially, this has produced

significant successes in the management of numerous lepidopteran and coleopteran

crop pests, particularly in reducing costs and improving yields. This has yet to be

deployed as a strategy in citrus.

Scientists at the USDA-USHRL (United States Department of Agriculture-

Horticultural Research Laboratory) in Fort Pierce, FL identified a gene encoding a Bt

toxin (Cyt2Ca1) active against D. abbreviatus and engineered citrus to express it.

Evaluation of Cyt2Ca1expressing citrus is carried out in my study.

Other strategies for insect control through transgenic crop development include

the use of protease inhibitors that disrupt the insect digestive process, and double-

stranded RNA to interfere with gene expression. Protease inhibitors have shown

promise because many plants use naturally produced protease inhibitors as a strategy

to protect themselves from insects. However, in practice, the success of this strategy

has been limited because many insect pests, especially coleopterans and

lepidopterans, have multiple alternative digestive proteases that are over-expressed to

compensate for inhibited proteases. In mosquitoes, blocking digestive proteases was

shown to be a functional strategy when the mechanism that regulates the expression of

digestive trypsin was targeted (Borovsky and Mahmood, 1995; Borovsky and Meola,

2004). Feeding of soybean trypsin inhibitor and Aedes aegypti trypsin modulating

oostatic factor affected trypsin activity and trypsin biosynthesis, respectively in

Diaprepes abbreviatus (Yan et al., 1999). A similar approach is sought to control D.

21

abbreviatus with inhibitors of digestive cathepsin L. Since D. abbreviatus, utilizes both

trypsin and cathepsin for digestion (Yan et al., 1999), targeting both enzymes in concert

with their respective inhibitors may be effective at precluding protein digestion in the

weevil.

RNA interference (RNAi) can be harnessed for crop protection against pests.

RNAi is a gene regulatory mechanism at the mRNA level. Therefore, by down-

regulating the expression of essential genes, growth and development of insect pests

can be impaired, possibly leading to death. The adaptation of RNAi as a pest control

tool is nascent, but precocious. The first application of RNAi for pest control was

demonstrated in Coleoptera and Lepidoptera (Baum et al., 2007; Mao et al., 2007). In

the short time since, RNAi applications has expanded exponentially into insect orders of

agricultural and medical importance (Coutinho-Abreu et al., 2010; Huvenne and

Smagghe, 2010). Novel attempts for D. abbreviatus control may benefit from RNAi.

In summary, this thesis presents alternative strategies for D. abbreviatus

management by targeting digestive proteases with inhibitors and the silencing of gene

expression via RNA interference (RNAi) with double-stranded RNA sequences

(dsRNA). An attempt at making a broad spectrum inhibitor of proteases is pursued and

a preliminary study to identify gene targets for silencing are conducted. The Bt toxin,

protease inhibitors and dsRNA for RNAi are explored for a complimentary method, in a

concerted multiple target effort, to produce transgenic citrus rootstock expressing more

than one toxin with different modes of action to offset the selection of resistance in D.

abbreviatus.

22

Current Control of Diaprepes abbreviatus

In citriculture, no satisfactory control for D. abbreviatus is available. Heavy

reliance on chemical insecticides is the mainstay of management of D. abbreviatus.

This is also true in the West Indies, where the impact of natural enemies and predators

on populations of D. abbreviatus has not kept the pest below economic thresholds (Hall

et al., 2001). Dependence on chemical insecticides is not only deleterious to the

environment, toxic to beneficial and non-target insects, and harmful to human and

animal health, but the use of ovicides and adulticides is considered ineffective and/or

expensive (Lapointe, 2000).

Control strategies against D. abbreviatus have been deployed target the various

life stages of the insect: eggs that are vulnerable to parasitoids, larvae that dwell below

ground and adults that feed on foliage. Biological and chemical methods are the

contemporary methods for control. However, none of these strategies have proven to be

efficacious or cost-effective.

Cultural Control: Sanitary-based containment and quarantine measures have

been practiced as a means of limiting the spread of D. abbreviatus. Since the insect

prefers to “hitchhike” as adults on plants and larvae in soil transported by people, this

has by some measure limited the spread of D. abbreviatus, though this is quite difficult

to ensure (Weissling et al., 2009).

Biological Control: Several egg parasitoids have been released in the field for

pest suppression: Quadrastichus haitiensis Gahan (Hymenoptera: Eulophidae),

Ceratogramma etiennei Delvare (Hymenoptera: Trichogrammatidae), Aprostocetus

vaquitarum Wolcott (Hymenoptera: Eulophidae), Fidiobia dominica Evans and Peña

(Hymenoptera: Platygastridae), and Haeckeliania sperata Pinto (Hymenoptera:

23

Trichogrammatidae), (Hall et al., 2001; Ulmer et al., 2006). These parasitic wasps have

been introduced to Florida from the West Indies. Results of releases of Q. haitiensis and

A. vaquitarum have not been spectacular. The wasps have failed to spread and expand

their range, perhaps as a result of concurrent insecticide use, and therefore have not

had the desired effect. Another possible explanation is the climatic differences between

the geographic origin of these wasps and Florida. Quadrastichus haitiensis, C. etiennei,

and A. vaquitarum are reported to parasitize greater than 35% of eggs in southeastern

Florida, though it is not clear what levels of control they are achieving statewide.

The entomopathogenic nematode Steinernema riobravis Cabanillas has also

been used in the field for D. abbreviatus control (Stuart et al., 2004). The nematode is

applied to the soil through the irrigation system. On reaching the weevil larvae, it

penetrates into the body and releases mutualistic bacteria that multiply and cause

septicemia. The nematode larvae feed on the decomposing cadaver, grow, and as

adults produce offspring that leave the host cadaver and infect other larvae in the

vicinity. Grower acceptance of the product to augment IPM strategies was mixed

(Georgis et al., 2006) and, due to problems with assuring proper quality control

associated with shipment and storage conditions, this product is no longer

recommended (Weissling et al., 2009). The nematodes used as biological control

agents are constrained by their requirements for efficacy that include optimal

temperature, appropriate soil type, and adequate irrigation (Shapiro-Ilan et al., 2006).

Chemical Control: The primary method of D. abbreviatus control by growers

has been insecticides targeted at both larvae and adults. Since the most damaging

stage of the insect is subterranean, this method has also been met with difficulty due to

24

the inability to reach the larvae with the insecticides in toxic doses. Insecticides that

have been used recently include bifenthrin and carbaryl to control adults and

imidacloprid against larvae (Jetter and Godfrey, 2009). Chlorinated hydrocarbons -

heptachlor, dieldrin, and chlordane - were the first insecticides used against the pest

(Nigg et al., 1999) but discontinued for environmental impact concerns.

Other methods listed include the use of oils to prevent leaves from sticking

together, thus exposing eggs to desiccation and also making them vulnerable to

predators and parasitoids. Desiccation of eggs reduces the hatchability of eggs and can

help bring down the population of the weevil in the field.

Future Control of Diaprepes abbreviatus

It is evident that no safe and effective control method is available for this pest.

Dependence on insecticides is inefficient because it is difficult to reach the root system

where larvae cause the most damage and parasitoids have failed to establish and

spread. Novel, cost-effective, and efficient control is needed for D. abbreviatus. For

these reasons, the use of transgenic technology is a reasonable and perhaps the only

strategy to provide control of this weevil.

Pests are a major obstacle to large scale agriculture and consequently their

management has always been important. When properly deployed, the transgenic

technology is efficacious. In the 25 years since the first deployment of transgenic crops,

no adverse effects have been realized (Romeis et al., 2013) and with respect to insect

control, reductions in the need for chemical pesticides have been touted (Lu et al.,

2012). Since the initial introduction of commercially available transgenic crops in 1996,

there are now an estimated 170.3 million planted hectares, up from 1.7 million hectares

in 1996 (James, 2012). Prominent among transgenic plants are insect resistant

25

varieties: Bt-engineered sweet corn to protect against the corn earworm, Helicoverpa

zea Boddie (Tabashnik et al., 2008), against the pink bollworm (Pectinophora

gossypiella Saunders) in cotton (Kathage and Qaim, 2012; Tabashnik et al., 2012), and

against the Colorado potato beetle (Leptinotarsa decemlineata Say) in potato (Zhou et

al., 2012).

A major challenge in pest management is the development of resistance to

control strategies. Although transgenic crop varieties have surpassed predictions for the

onset of resistance (Bates et al., 2005), reports of field-evolved resistance in insects

(Tabashnik, 2008; Liu et al., 2010; Gassmann et al., 2011) are beginning to emerge

after more than a decade of commercial availability.

The first commercially available transgenic plants were engineered to express

single toxin genes. Insect management strategies, such as the inclusion of refuges

(localized plantings of non-transgenic crops than can function as a location where the

pest insect can reproduce where they are not under selection pressure), have played a

big role in forestalling resistance but there is concern that this practice is only slowing

the development of resistance and not preventing it. Fortunately, the inherent

capabilities of transgenic plants can be enhanced. In a bid to delay - better still,

preclude - the inception of resistance, it has been advocated that future transgenic

plants be developed to express multiple toxins. This is known as the pyramiding or

stacking strategy (Roush, 1997). This is the strategy suggested for transgenic citrus.

Transgenic Control of Diaprepes abbreviatus

Production of transgenic citrus plants expressing multiple deleterious toxins

against D. abbreviatus is an attractive and feasible control strategy. The key is to

identify multiple physiological/biochemical processes within the weevil that can be

26

targeted as interdiction points. The weevil’s digestive process presents an excellent

target for interdiction. This can be achieved by disrupting gut integrity and/or inhibiting

protein digestion, and interfering with the expression and consequent function of

important genes. Toxins with different modes of action will be particularly advantageous

in a multi-pronged attack on the insect to delay the onset of resistance. Several

candidate toxins at various stages of evaluation are reported herein.

Candidate Toxins

This research is based on the anticipation of imminent resistance to a single-

toxin transgenic strategy for D. abbreviatus. The identification and evaluation of toxins

that work in a variety of ways is given priority. Ideally, toxins with unrelated modes of

action are most suitable to reduce the probability of selecting for individuals with a

particular set of resistant genes. Specifically, the strategies to control D. abbreviatus in

a transgenic citrus variety include the Bt toxin Cyt2Ca1 to disrupt gut integrity.

Secondly, cysteine protease inhibitors (CPI) that target an important digestive enzyme,

cathepsin L1, are also investigated. Finally, double-stranded RNA sequences/molecules

to interfere with gene expression via mediated RNA interference are also screened.

Endotoxins from Bacillus thuringiensis. Insecticidal proteins (endotoxins)

produced by Bt, a soil microorganism, have been widely used in genetic engineering of

food crops in the past two decades. However, the bacterium’s use as an insecticidal

agent extends beyond a century (Crickmore, 2006) as foliar sprays. They are target-

specific with activities against different orders of insects - Lepidoptera, Diptera, and

Coleoptera. The insecticidal proteins, Crystal (Cry) and Cytolytic (Cyt) proteins are

produced during the sporulation phase of the bacteria. Bacillus thuringiensis strains

have evolved these toxins to colonize insects because they serve as a ready source of

27

nutrition for germination of the dormant spore, though a small minority of Cry toxins are

nematicidal (Marroquin et al., 2000; De Maagd et al., 2001). The Cry proteins are

secreted as protoxins by the bacterium, and after ingestion, are activated by gut

proteases to active toxins. The activated toxins bind to receptors in the gut epithelium,

resulting in the creation of pores and lyses of midgut epithelial cells (De Maagd et al.,

2001). The Cyt toxins have a different mechanism of action, but are believed to work in

a detergent-like mechanism by interacting directly with membrane lipids to insert into

the membrane (Manceva et al., 2005). In the Cry engineered transgenic plants, they are

engineered to express the active forms of the enzymes (precluding the need for

activation) and optimized with plant codon usage (Vaughn et al., 2005), making them

more robust and versatile.

Bacillus thuringiensis strains were screened to identify toxins active against D.

abbreviatus (Weathersbee et al., 2006). From this work, a candidate Cyt toxin,

Cyt2Ca1, was identified with activity against D. abbreviatus larvae. A synthetic Cyt2Ca1

gene was constructed with codon optimization specific for highly expressed genes in

citrus and this synthetic gene was engineered into alfalfa in a proof of concept approach

(alfalfa supports all life stages of the weevil). Preliminary work with transgenic alfalfa

(Figure 1-1) has shown that this is a promising strategy for D. abbreviatus control, and

therefore, Cyt2Ca1 was subsequently engineered into citrus.

28

Figure 1-1. General plant and root damage as a result of Diaprepes abbreviates feeding on transgenic versus wild-type alfalfa (Shatters, unpublished data). Four weeks after the addition of 10 larvae to individually potted plants, it is evident that transgenic plants had greater resistance to D. abbreviates injury (Photos by R.G. Shatters, Jr.).

Peptide inhibitors against protein digestion. The ability to break down

proteins and complex carbohydrates, by digestive proteases in the gut, into free amino

acids and simple sugars is essential for growth and development of insects (Bown et al.,

2004). Therefore, the inhibition of digestive proteolysis by protease inhibitors has been

suggested as a pest management strategy (Haq et al., 2004; Christou et al., 2006;

Schlüter et al., 2010).

In many coleopteran species such as the cowpea bruchid (Callosobruchus

maculatus Fabricius) and the western corn rootworm (Diabrotica virgifera virgifera

LeConte), cysteine proteases are the major digestive enzymes (Koiwa et al., 2000; Zhu-

Salzman et al., 2003). In D. abbreviatus, cathepsins (cysteine proteases) are one of two

important groups of enzymes responsible for protein digestion and predominantly active

under acidic conditions. The other is a serine protease with trypsin-like activity that

Cyt2Ca1 Alfalfa

Cyt2Ca1 Alfalfa

Wild-type Alfalfa Wild-type Alfalfa

29

functions in the alkaline range (Yan et al., 1999). Targeting the digestive proteases in

the midgut of D. abbreviatus may contribute to a pyramided strategy.

Preliminary studies have been done to determine the effect of an inhibitor, trypsin

modulating oostatic factor (TMOF), on D. abbreviatus trypsin. This inhibitor is a

decapeptide (YDPAPPPPPP) that has been shown to be the physiological signal that

terminates trypsin biosynthesis through translational control in mosquitoes and other

insects (Borovsky et al., 1993, 1995, 1996). In vitro, TMOF has been shown to inhibit

the activity of gut trypsins, and feeding TMOF to larval D. abbreviatus reduces weight

gain and results in reduced gut trypsin activity (Yan et al., 1999). Dov Borovsky and

Robert Shatters (personal communication) have cloned and expressed TMOF in alfalfa

and demonstrated that feeding citrus weevil the transgenic alfalfa leaves inhibited

trypsin activity in the gut. However, because D. abbreviatus also relies on cathepsins for

digestive protein degradation, blocking trypsin alone is not sufficient to induce significant

mortality. Therefore, research was conducted to characterize the nature of D.

abbreviatus digestive tract cathepsins and to test a strategy to block their activity that

could be deployed through transgenic means.

Cathepsins belong to the papain family of proteins, and many are lysosomal

proteins (Cristofoletti et al., 2005) involved in cellular protein catabolism (Kamboj et al.,

1993; Stachowiak et al., 2004). Through numerous studies of cathepsins, several

classes have been identified (Kirschke, et al., 1977; Turk et al., 2000) and classified

based on characteristics such as their activities with unique substrates and specific

inhibitors, and pH optimum for activity (Rawlings et al., 2012). In humans, mutations

affecting cathepsin activity are implicated in serious diseases such as arthritis, muscular

30

dystrophy and tumors (Hasnain et al., 1992; Gewies and Grim, 2003). In insects,

cathepsins are involved in digestion, embryogenesis, molting, tissue remodeling, and

reproduction (Bown et al., 2004; Deraison et al., 2004; Pyati et al., 2009).

In members of Coleoptera and Hemiptera known to utilize digestive cathepsins

(Terra, 1990), the digestive cathepsins are secreted into the gut lumen by the epithelial

cells. Cathepsins are synthesized as preproproteins; a polypeptide with three

identifiable regions: a pre-region which serves as a signal peptide that causes the

enzyme to enter the cellular secretory pathway where it is ultimately secreted into the

gut lumen; a pro-region which is essential for protein folding and regulates enzymatic

activity; and must be cleaved to generate the mature and active protein (third region of

the preproprotein). All cathepsins are characterized by a cysteine residue at the active

site (Turk et al., 2000) with a broad specificity for cleaving peptide bonds. Upon

synthesis, they remain inactive until the prepro- region of the polypeptide is removed.

The N-terminal pre-region is co-translationally removed (Philip et al., 2007), whereas

the pro-region is cleaved via autocatalysis at the optimum pH of the mature protein. The

pro-region, approximately 11-13 KDa in size, regulates enzymatic activity of the

proteinase domain by its interaction with the active site to block proteolytic activity (Tao

et al., 1994 and Philip et al., 2007). The pro-region is cleaved via autocatalysis at the

optimum pH and the molecule is activated.

More recently, genes encoding small peptides (approximately 11 KDa) with

significant homology to the pro-region of cysteine proteases have been characterized in

the digestive tract of insects (Kurata et al., 2001; Deshapriya et al., 2007; Miyaji et al.,

2010). The activities of the digestive cathepsins are modulated by these smaller

31

proteins (Figure 1-2). Two sequences encoding D. abbreviatus cathepsin L protease

inhibitors (CPI 1 and CPI 2) synthesized in the gut epithelial cells have recently been

discovered (Borovsky, unpublished results). CPI 1 is secreted into the gut lumen;

however, CPI 2 lacks a signal peptide and is believed to function in the epithelium

cytoplasm. Thus, it was hypothesized that the CPI1 modulates cathepsin activity in the

gut lumen. My thesis reveals research conducted to characterize the activity of CPI1

and attempts to modulate its inhibitor activity, to be utilized as a toxin expressed in

citrus to confer D. abbreviatus mortality.

32

CPI 2 ---------- -----MSAPT -KAP--SYLS DQEEWEKFKT GFNRNYDSSD

CPI 1 MLVKVFLLVV LAAVAMSAPS DTAPKQKSLS VEEHWNNFKT KFNRNYESPE

Cathepsin L 1 ---------- MKVFIAACLL VAVSATVLEE TGVKFQAFKL KHGKTYKNQV

Cathepsin L 2 -------MYS LVVLLATLVA YSHAISYQVL VQEQWEQFKL EHGKVYESES

CPI 2 EEAKRFNIFQ QNLQSIREHN EKFERGETTF TQGINQFTDL TKEEFKARHT

CPI 1 EESKRFEIFK NNLKDIQAHQ KKYEAGEVSY QQGVNDFTDL THEEFLATHT

Cathepsin L 1 EETARFNIFK DNLRAIEQHN VLYEQGLVSY KKGINRFTDM TQEEFRAFLT

Cathepsin L 2 ENEYRQSVFM ENLFQINEHN KLYEMGLSSY QMAMNHLGDL TKDEFMRIYT

CPI 2 GLLRRPPQE- ---------- ---------- ---------- ----------

CPI 1 MHFNPKPKS- ---------- ---------- ---------- ----------

Cathepsin L 1 LSSSKKP--- HFNTTEHVLT G--------- ---------L AVPDSIDWRT

Cathepsin L 2 VNMPQLPQSE NLSDSEPWLD LPQDLQGFVT YALPTNLDEV DLPTDIDWRQ

CPI 2 ---------- ---------- ---------- ---------- ----------

CPI 1 ---------- ---------- ---------- ---------- ----------

Cathepsin L 1 KGQVTGVKDQ GNCGSCWAFS VTGSTEAAYY RKAGKLVSLS EQQLVDCS-T

Cathepsin L 2 KGAVTPVKNQ RNCGSCWSFS ATGALEAQWF KKTNKLISLS EQQLVDCSGR

CPI 2 ---------- ---------- ---------- ---------- ----------

CPI 1 ---------- ---------- ---------- ---------- ----------

Cathepsin L 1 DINAGCNGGY LDETFTYVKS KG-LEAESTY PYKGTDGSCK YSASKVVTKV

Cathepsin L 2 YGNHGCHGGW MHWAFGYIKE NGGIDTEQSY PYTAKDGRCA YKPGNKAATV

CPI 2 ---------- ---------- ---------- ---------- ----------

CPI 1 ---------- ---------- ---------- ---------- ----------

Cathepsin L 1 SGHKSLKSED ENALLDAVGN VGPVSVAIDA TY-LSSYESG IYEDDWCSPS

Cathepsin L 2 S-QVIMVPRG ENQLAAKVSS VGPISIAAEV SHKFQFYHSG VYDEPQCGHS

CPI 2 ---------- ---------- ---------- ---------- ----------

CPI 1 ---------- ---------- ---------- ---------- ----------

Cathepsin L 1 ELNHGVLVVG YGTSNGKKYW IVKNSWGGSF GESGYFRLLR GKN-ECGVAE

Cathepsin L 2 -LNHAMLAVG YGSMGGKNFW LVKNSWGTGW GDQGYIRMAK DKNNQCGIAL

CPI 2 -------

CPI 1 -------

Cathepsin L 1 DTVYP--

Cathepsin L 2 MASYPGV

Figure 1-2. Amino acid sequence alignments of cathepsin L protease inhibitors (CPI 1 and 2) to cathepsin L 1 and 2 in Diaprepes abbreviatus. The pre-region of the cathepsin L 1 sequence in highlighted in green and the pro-region is highlighted in yellow.

Double-stranded RNA. RNA interference (RNAi) is an innate mechanism of

gene regulation in eukaryotes. It is an evolutionarily conserved gene-regulatory pathway

(Meister and Tuschl, 2004) that follows several different pathways ultimately to the

33

same fate - the silencing of genes. RNAi was known as a process in plants for many

years prior to its discovery in animals about a decade ago in the nematode

Caenorhabditis elegans (Maupas) (Fire et al., 1998). It has since been identified as a

potential mechanism for pest/vector control. By down-regulating essential gene function

(Price and Gatehouse, 2008), growth, development, and reproduction can be

hampered, thus leading to death or less fit individuals.

The underlying theme among all the different pathways is the requirement of a

double-stranded RNA template. Templates may be of varying length that specifically

hybridize to and result in the cleavage of a specific mRNA sequence transcribed from a

gene. Principally, dsRNA templates are processed by specific endonucleases,

Dicer/Drosha, into short RNA duplexes. These duplexes of short-interfering RNA

(siRNA) are typically 20-23 base pairs in length and staggered. They mediate

translational repression and mRNA degradation via the catalytic activity of argonaute

proteins, as part of a silencing complex (RNA-induced silencing complex: RISC) that

contains single-stranded siRNA (Meister and Tuschl, 2004). The single-stranded siRNA

recognizes complementary strands that are subsequently destroyed.

RNAi in insects is believed to constitute a part of an immune response to viral

infection (Keene et al., 2004). It is possible to manipulate an insect’s RNAi machinery to

attack itself, by presenting dsRNA complementary to its genes, such that it down

regulates the expression of these genes. A limitation of the RNAi mechanism exists in

insects, however. Contrary to C. elegans, plants, and fungi, there is no apparent

amplification of the dsRNA template in insects. There is yet no evidence of the presence

of the enzyme RNA-dependent RNA polymerase (RdRP) in insects for amplifying the

34

RNA signal (Tomoyasu et al., 2008). Despite this limitation, it is still possible to induce

an RNAi response in insects because of the systemic spread of dsRNA in insects

(Huvenne and Smagghe, 2010).

Gene silencing via RNAi has therefore been identified as a potential mechanism

for pest/vector control (Gordon and Waterhouse, 2007). Work on RNAi on the western

corn rootworm and the Colorado potato beetle shows that ingestible dsRNA may be

used in the future to control insects as transgenes expressed in transgenic plants

(Baum et al., 2007; Mao et al., 2007). Numerous other examples (Price and Gatehouse,

2008) highlight the potential of RNAi for pest control and would place it on the frontier of

pest management technology.

The Watson-Crick hybridization requirement of the RNAi mechanism makes it a

very valuable tool for future pest control. Thus, RNAi can be designed to be very

specific (Whyard et al., 2009). However, off-target effects can also occur (Nunes et al.,

2013). Additionally, lessons gleaned from the Shatters laboratory at the USDA-USHRL

investigating RNAi for control of the Asian citrus psyllid, Diaphorina citri Kuwayama,

indicate that it is possible to induce mortality with significant doses dsRNA of any nature

or source. Therefore, a dose-response effect is vital in RNAi experiments.

Bacillus thuringiensis Cyt2Ca1, peptide inhibitors, and dsRNA sequences will

provide a three-pronged approach to D. abbreviatus control in a transgenic strategy.

The efficacy of each of these approaches needs to be independently verified before any

future synergistic studies. My thesis sets out to study these approaches separately. A

discussion of delivery modes in a transgenic strategy is necessary, and I express an

opinion about this in the following section.

35

Delivery of Toxins in a Transgenic Strategy

Having identified suitable toxins for control through artificial diet bioassays, it is

important then to consider a variety of delivery methods for a transgenic variety.

Initially, transgenic alfalfa expressing identified D. abbreviatus toxins was promoted as a

shield crop in citrus groves because the weevil also eats alfalfa. In a proof-of-concept

experiment, the laboratories of Dov Borovsky and Charles Powell at the University of

Florida and Robert Shatters in the USDA-USHRL cloned TMOF active against D.

abbreviatus in alfalfa. Feeding citrus weevil the transgenic alfalfa leaves stopped

trypsin activity in the gut of the weevil (Borovsky et al., unpublished data). On the other

hand, avoidance of transgenic alfalfa and a preference for citrus may compromise the

success of such a strategy and possibly exacerbate the pest status of the weevil.

Citrus is attractive for engineering against D. abbreviatus because of the way it is

cultivated. Commercial cultivation of citrus uses trees produced by grafting the desired

fruiting cultivars (scions) onto rootstocks selected for traits such as resistance to insects

and diseases. Citrus rootstocks can thus be engineered for resistance to D.

abbreviatus, but not the scions, leaving fruits without transgenic genes. The fact that

recombinant proteins would not be expressed in the fruits may boost consumer

acceptance, as the ultimate goal of this project is to have rootstock strains that will

provide a commercially viable pest management strategy in citriculture. An engineered

citrus rootstock delivery mechanism has the principal advantage of being target-specific

on two counts. Firstly, the range of insects that could be affected is narrowed down to

only insects that directly feed on citrus, excluding insects that do not feed on citrus.

Secondly, but more importantly, these toxins are sourced directly from D. abbreviatus

specifically.

36

A third delivery mechanism is to use citrus tristeza virus (CTV) as a vector of

toxins. Citrus tristeza virus, an infectious agent commonly present in citrus, also lends

itself as a vector of dsRNA and protein molecules in citrus (Folimonov et al., 2007).

Engineered to be infective but not pathogenic, several CTV-based transient-expression

vectors have been developed for gene expression or silencing studies in, but not limited

to, citrus (Dawson et al., 2010). The CTV vector has been designed for systemic field

infection of trees to express copious amounts of recombinant gene product for up to

several years. As the replicative form of CTV is double-stranded RNA (Tatineni et al.,

2010), it also enables manipulation for the expression of dsRNA molecules, without a

hair-pin loop targeting specific organisms in an RNA-mediated interference mechanism.

Thus it will be possible to employ the CTV-vector to express peptide inhibitors and/or

dsRNA toxins identified for specific D. abbreviatus control under field conditions. The

greatest advantage of using CTV as a delivery tool is that it drastically cuts the time

typical for the development of resistant trees. Importantly, it removes the requirement to

completely re-plant citrus groves with transgenic strains as it is possible to infect

existing trees with the engineered virus.

37

CHAPTER 2 RECOMBINANT EXPRESSION AND CHARACTERIZATION OF DIAPREPES

ABBREVIATUS CATHEPSIN L1 AND THE ROLE OF CATHEPSIN L PROTEASE INHIBITORS IN THE REGULATION OF CATHEPSINS INVOLVED IN DIGESTION

Proteins are an important food source for D. abbreviatus, similar to many

organisms. The literature is replete with many examples where the inhibition of digestive

proteases resulted in death and retardation of growth of many pest insects (Oppert et

al., 1993; Macedo et al., 2003; Vila et al., 2005). Therefore, inhibiting protein digestion

in insects with protease inhibitors (PIs) has been suggested for pest control (Bown et

al., 2004; Christou et al., 2006). The digestive proteases of targeted pest insects require

in depth characterization to find potential inhibitors of their activities. Cathepsins

constitute one component of a dual proteolytic system in D. abbreviatus (Yan et al.,

1999). Diaprepes abbreviatus cathepsins are not as well characterized as the trypsins

(the other component). Due to the role of D. abbreviatus cathepsin in digestion, they

could be an important target for inhibition.

Gut transcriptome studies from larval stages of D. abbreviatus identified

numerous cathepsins that are turned on or off depending on the life-stage and age of

the insect, suggesting their need for growth and development (Shatters, unpublished

data). Additionally, crude D. abbreviatus gut assays indicate the predominance of

cathepsins active in the acidic range between pH 4-6.5 in a midgut that is partitioned

into acidic (anterior) and basic (posterior) regions, similar to the aphid, Aphis gossypii

Glover (Deraison et al., 2004), and the weevil Otiorhynchus sulcatus (Fabricius)

(Edwards et al., 2010). In order to inhibit the enzyme in a possible transgenic control

strategy of D. abbreviatus in citrus, pure recombinant cathepsins will be invaluable for

kinetic studies. Two of the highly expressed cathepsin L homologues were selected for

38

recombinant synthesis however, only cathepsin L1 (DaCatL1) was successfully

expressed and purified. The results of the characterization of DaCatL1 are presented

herein.

In vitro enzymatic assays used to characterize cathepsins feature small peptide

substrates consisting of a sequence of amino acid residues linked to a measurable

leaving group. Also, Positional Scanning Synthetic Combinatorial Libraries, enable the

determination of the preferred amino acid sequences of substrates for various

proteases including cathepsin L (Cotrin et al., 2004; Choe et al., 2006).The leaving

groups of the substrates may be fluorogenic or chromogenic (Tchoupé et al., 1991). The

classes of cathepsins are grouped according to chemical properties available in the

MEROPS database including the sequence of amino acids in suitably cleaved

substrates (Berger and Schechter, 1970; Rawlings et al., 2012). Cathepsin L has been

shown to have substrate tolerance for hydrophobic residues in the P2 and P3 positions

(Otto and Schirmeister, 1997) with a preference for leucine or arginine at the P2 position

(Clara et al., 2011).

The activities of cathepsins are regulated by small peptide inhibitors called

cystatins. Two endogenous regulators of D. abbreviatus digestive cathepsins, cathepsin

L protease inhibitors (CPI 1 and 2), have been identified (Borovsky, Shatters, and

Powell, unpublished data). Their inhibition of cathepsins is competitive and reversible

and thus, it may be possible to modulate this interaction. Amino acids in the CPI

molecule are predicted to be important for their inhibitory activity. Therefore amino acid

substitutions in CPI may lead to a mutant with modified chemical interaction with

cathepsin. Cathepsins catalytic cysteine is hypothesized to form a disulfide bridge with a

39

proximate cysteine on a CPI mutant. The disulfide bridge formed between the mutant

inhibitor, and enzyme would make the mutant a better inhibitor due to the irreversible

interaction. Site-directed mutagenesis is a technique to introduce amino acid

substitutions in proteins. Polymerase chain reaction utilizing primers that incorporate

codon change flanked by long complementary sequences is used to generate mutations

in a cDNA template via the technique.

Although no PIs are available commercially, as ready-to-use formulations or in

transgenic cells, they can still be complementary to conventional approaches using

chemical and biological pesticides in future pest control strategies. They can make

target pests more vulnerable to other toxins, such as Bt toxins or dsRNA, in a stacked

transgenic approach. Thus, DaCatL1-CPI interactions need to be verified and exploited.

Materials and Methods

Molecular Modeling and Docking of Cathepsin L Protease Inhibitor to D. abbreviatus Cathepsin L1

Homology modelling of DaCatL1 was done using the YASARA Structure program

(Krieger et al., 2002). Different models of DaCatL1 were built from the X-ray coordinates

of the procathepsin L1 from Fasciola hepatica Linnaeus (PDB code 2O6X) (Stack et al.,

2008), the larval midgut procathepsin L3 of Tenebrio molitor Linnaeus (PDB code

3QT4) (Beton et al., 2012), the Cys25Ala mutant of human procathepsin S (PDB code

2C0Y) (Kaulmann et al., 2006), the human procathepsin L (PDB Code 1CS8)

(Coulombe et al., 1996), and the activated cathepsin L of Toxoplasma gondii Nicolle

and Manceaux in combination with the propeptide (PDB code 3F75) (Larson et al.,

2009). Finally, a hybrid model of DaCatL1 was built up from the five previous models.

40

The CPI1 was similarly modelled using the X-ray coordinates of the propeptide

from the larval midgut procathepsin L2 from T. molitor (PDB code 3QJ3) (Beton et al.,

2012), the procathepsin L1 from F. hepatica (PDB code 2O6X) (Stack et al., 2008), the

larval midgut procathepsin L3 of T. molitor (PDB code 3QT4) (Beton et al., 2012), the

Cys25Ala mutant of human procathepsin S (PDB code 2C0Y) (Kaulmann et al., 2006),

and the activated cathepsin L of T. gondii in combination with the propeptide (PDB code

3F75) (Larson et al., 2009). A hybrid model for CPI1 was built from the five previous

models.

PROCHECK (Laskowski et al., 1993) was used to assess the geometric quality

of the three-dimensional models. In this respect, all of the residues of DaCatL1 were

correctly assigned in the allowed regions of the Ramachandran plot except for 9 Gly

residues which occur in the non-allowed region of the plot. Using ANOLEA (Melo and

Feytmans, 1998) to evaluate the model, only 12 residues of DaCatL1 out of 306

exhibited energy over the threshold value. All residues are located in the loop region

connecting alpha-helices and beta-sheets. Molecular models were drawn with the

UCSF Chimera package (Pettersen et al., 2004).

The T. gondii cathepsin L (TgCPL) in complex with its propeptide (PDB code

3F75) (Larson et al., 2009) was taken as a model for docking experiments of the

cathepsin L inhibitor CPI1 on activated DaCatL1. Docking pattern was also rendered

with Chimera.

The SDSC Biology Workbench (San Diego, CA) was used for amino acid

sequence comparisons between DaCatL1 precursor (GenBank: ABG73217).1) and the

Homo sapiens cathepsins L1 isoform 1 preproprotein (NCBI Reference Sequence:

41

NP_001903.1), K (GenBank: AAH16058.1), S (GenBank: AAC37592.1), B (GenBank:

AAH95408.1), and H (GenBank: EAW99143.1).

Rearing of Diaprepes abbreviatus

Diaprepes abbreviatus larvae were obtained from a colony maintained at the

USDA-USHRL, Fort Pierce, Florida on an artificial diet described by Lapointe and

Shapiro (1999).

Bacterial Strains and Plasmids

Escherichia coli strains OneShot TOP 10F’ and DH5-alpha (Invitrogen, Grand

Island, NY) were used for sub-cloning, purification, and sequencing of plasmid DNA.

The plasmid used for checking DNA sequences was pCRTM2.1 (Invitrogen). Expression

plasmids used were pET DUET (EMD Millipore, Billerica, MA) for cathepsin L protease

inhibitors (both wild-type and mutant forms) and pET-41a (+) (EMD Millipore) for

DaCatL1. These plasmids permit the expression of proteins as fusion products, with the

incorporation of a polyhistidine tag (both) and a glutathione-S transferase tag (pET-

41a(+)). A plasmid purification kit (Qiagen, Valencia, CA) was used for all purifications

according to manufacturer’s protocol. An IPTG (Isopropyl β-D-1-thiogalactopyranoside

)-inducible derivative of E. coli strain BL21 (DE3), Rosetta™ (DE3) pLysS Competent

Cells, purchased from EMD Millipore, was used for expression of recombinant proteins.

Bacterial strains were grown in Luria broth (LB) media at 37 °C except after inducing

with IPTG for protein synthesis and expression at which time cells were grown at 25 °C

for at least 4 h up to overnight.

Primers

The primer sequences (5’-3’) listed below were obtained from IDT® (San Diego,

CA) and used for the cloning of DaCatL1, GenBank accession number: GenBank:

42

DQ667143.1 (GstproCatL1-f and GsrproCL1-r) and site-directed mutagenesis of CPI1,

GenBank accession number: EU009453.1, to make the mutant CPI (CPI1mut-f and

CPI1mut-r).

GstproCL1-f:

CTCAATCACTAGTACCGTCTTGGAGGAGACAGGTGTCAAAT

GstproCL1-r:

GCAGTCACTCGAGTATAATTGGATATACGGTATCTTCAGCAACTCC

CPI1mut-f:

GAATTTCTTGCTACTCACACGTGTCACTTCAATCCCAAACCCAAG

CPI1mut-r:

CTTGGGTTTGGGATTGAAGTGACACGTGTGAGTAGCAAGAAATTC

Cathepsin L 1 and Cathepsin L Protease Inhibitor Constructs

The DNA sequences of the pro- and active regions of DaCatL1 were amplified

from a cDNA library and cloned into the expression vector pET-41a (+) DNA. The

primers GstproCL1-f (incorporating the cut site for SpeI) and GstproCL1-r (incorporating

the cut site for XhoI) were used to amplify DaCatL1 using standard polymerase chain

reaction (PCR) protocols. The following thermal cycling conditions were used: 95 °C for

10 min, followed by 30 cycles of 95 °C for 15 sec, 57 °C for 30 sec and 72 °C for 30 sec,

and a final extension at 72 °C for 10 min. The PCR product was verified on an agarose

gel via electrophoresis. The band holding the product was cut from the agarose gel. The

DaCatL1 fragment was recovered from the agarose gel band using Wizard® SV Gel and

PCR Clean-Up System (Promega, Madison, WI) following manufacturer’s instructions.

The restriction enzyme ends incorporated into the amplification primers were cut from

the DaCatL1 fragment with restriction enzymes (SpeI and XhoI, obtained from New

43

England Biolabs®, Ipswich, MA), and the fragment was then cloned into the expression

vector pET-41a. One Shot Top 10 (Invitrogen) competent E. coli cells were transformed

with pET-41a (100 ng) and grown on LB media containing kanamycin (50 µg/mL).

Single colonies were selected and grown overnight at 37 °C in LB media containing

antibiotic in a shaking incubator at 225 rpm. Plasmid (pET41-a) was then extracted from

the E. coli cells using a plasmid extraction kit (Qiagen) and digested with SpeI and Xho1

as described for the DaCatL1 fragment. The linearized vector was separated by

agarose gel electrophoresis, cut from the gel and cleaned as previously described.

SpeI/XhoI-digested pET 41a and DaCatL1 were ligated using T4 DNA ligase overnight

at 4 °C. The ligated product was transformed into One Shot Top 10 competent cells,

single colonies were selected and the plasmids extracted and sequenced. Several

colonies contained the DaCatL1 insert, and glycerol stocks were made and stored at

minus 80 °C.

Cathepsin L protease inhibitor (wild type) (CPI 1, GenBank: EU009453.1) was

cloned into the pETDuet-1 (EMD Millipore) at the multiple cloning site-1 between the

restriction sites BamHI and NotI for synthesis of the recombinant protein as a fusion

product with a Histidine tag to enable selective purification based on affinity to a Ni-NTA

resin.

DNA fragments were sequenced by the genomics core facility at the USDA-

USHRL, and University of Florida-Interdisciplinary Center for Biotechnology Research

(UF-ICBR), Gainesville.

Site-Directed Mutagenesis

Primers (CPI1mut-f and CPI1mut-r) were synthesized to replace Lysine 101 with

Cysteine residue in the CPI1 construct (CPI1 fragment cloned into the pETDuet-1

44

Vector (EMD Millipore) using the QuikChange II XL Site-Directed Mutagenesis Kit and

manufacturer protocol (Agilent, Santa Clara, CA). A PCR was set up with the CPI1

construct as a template, according to the manufacturer’s instructions. The amplified

product was digested with DpnI, cleaned using the Wizard® SV Gel and PCR Clean-Up

System (Promega), and cloned into XL10-Gold Ultracompetent Cells (Agilent). Single

colonies were selected from LB plates (100 µg/mL carbenicillin), grown in LB

carbenicillin media, and plasmids extracted using plasmid extraction kit from Qiagen.

Plasmids were sequenced as described below, and verified constructs were

transformed into Rosetta™ (DE3) pLysS Competent Cells (EMD Millipore) for protein

synthesis.

Expression of Recombinant Proteins

Rosetta™ (DE3) pLysS Competent Cells (EMD Millipore) cells were transformed

with pET-41a carrying DaCatL1 sequence. Multiple colonies with the engineered vector

were tested to verify the biosynthesis of the recombinant proteins. Cell cultures (5 mL)

were initially grown at 37 ºC, 225 rpm in LB medium (30 μg/mL chloramphenicol and 50

μg/mL kanamycin) to OD600 of 0.6. Isopropyl β-D-1-thiogalactopyranoside (IPTG) (1

mM) was then added to induce synthesis of recombinant proteins in treatment cultures.

Controls were grown similarly, except the IPTG induction step was omitted. The

temperature during the induction stage was maintained at 25 ºC for 4-6 h. Cells were

harvested by centrifugation at 16,000 rpm for 5 min at 4 ºC (Eppendorf's refrigerated

bench-top 5702 R centrifuge), re-suspended in 1 mL phosphate buffer saline (50 mM

NaH2PO4, 300 mM NaCl, pH 7.0), and added to 2-mL tubes containing 180 μm acid-

washed glass beads Cells were broken in a FastPrep Instrument for 1 min and

centrifuged at 14,000 rpm for 10 min at 4 ºC. The supernatant was collected and

45

aliquots (20 μL each) were run on sodium dodecyl sulfate polyacrylamide gel

electrophoresis (SDS-PAGE) using the 4-12% NuPAGE® Novex® Bis-Tris Gel (1.0 mm

thick) (Life Technologies, Carlsbad, CA). DaCatL1 fused with the GST tag ran as a

protein band of 60 KDa, whereas the CPI1 and CPI1mut ran as bands of 11 KDa. To

synthesize larger amounts of recombinant proteins, larger cultures of transformed E.

coli (4-6 L) were grown and harvested as before. Harvested E. coli cells were lysed

using a cell disruptor (Branson 450 Sonifier - SmithKline). However, the DaCatL1

quickly precipitated out of solution. To determine if adding surfactants/detergents

(obtained from Affymetrix, Santa Clara, CA) would improve solubility, the following

detergents were tested: Amphipol (20 ng/mL), CHAPS [3-((3-Cholamidopropyl)

dimethylammonium)-1-propanesulfonate] (8 mM), OGNG [Octyl Glucose Neopentyl

Glycol] (1 µM) and TritonTM X-100 (0.25% and 1%). Incorporating 1% TritonTM X-100 in

all buffers used from lysis stage through nickel resin (Ni-NTA, Qiagen) purification, and

size exclusion chromatography maintained the enzyme in solution. SDS-PAGE was

used to identify the recombinant protein followed by western blot analysis.

Protein quantification was done using the Quick Start™ Bradford (Bio-Rad,

Hercules, CA) method in all cases, except when proteins were solubilized in 1%

TritonTM-X100 in which the RC DC™ Protein Assay (Bio-Rad) was used. Synergy™ HT

Multi-Mode Microplate Reader (BioTek, Winooski, VT) was used to read absorbance.

Western blot analysis of recombinant proteins was done using the Xcell II blot

module from Invitrogen to transfer resolved proteins from SDS-PAGE onto a

nitrocellulose membrane. Antibodies raised against histidine tagged proteins- 6x-His

46

Epitope Tag Monoclonal Antibody kit (Thermo Scientific, Waltham, MA), was used to

locate lanes (bands) with DaCatL1, CPI1, and CPI1mut.

Western Lightning Plus ECL kit (PerkinElmer, Waltham, MA) was used to

visualize proteins on the nitrocellulose membrane with the aid of a KODAK imager

(Image Station 440CF).

Purification of Recombinant Proteins

After lyses of cells via sonicator, cell debris was removed by centrifugation

(Beckman XL-I analytical ultracentrifuge - 20,000 rpm at 4 °C for 40 min). Initial

purification of histidine-tagged recombinant proteins from the crude extract was done

using affinity resin (Ni-NTA Superflow -Qiagen). Equilibrated resin (5 mL; 50 mM Tris

buffer, pH 7, 300 mM NaCl, 1 mM BME, 1% TritonTM, 10% glycerol) was added to 50

mL crude extract and gently mixed in the cold for an hour. Unbound proteins were

removed by centrifugation at low speed (2,000 rpm, 5 min, 4 °C), and bound proteins

were washed with buffer (5 times) by mixing in the cold as before with 5 volumes of

buffer. Recombinant proteins were eluted from Ni-NTA resin with 250 mM imidazole in

buffer (above), by mixing as before. The Ni-NTA resin was removed by low speed

centrifugation and purified proteins were analyzed by SDS-PAGE and western blotting.

The recombinant proteins were then purified by size-exclusion chromatography

(HiprepTM 16/60 SephacrylTM S-100 HR) with an ÄKTA™FPLC™ system (GE

HealthCare Life Sciences, Piscataway Township, NJ). Eluted proteins were

concentrated on centrifugal filters (Amicon® Ultra-15 mL with 3000 and 10, 000 MW cut-

off for cysteine protease inhibitors and recombinant cathepsins respectively, EMD

Millipore), to each aliquot glycerol was added to a final concentration of 20%, and the

proteins were flash frozen and stored at -80 °C.

47

Enzymatic Assays

DaCatL1: The activity of DaCatL1 was initially assayed using three small

fluorogenic peptide substrates, Z-GPR-AMC (Z= benzyloxycarbonyl, GPR = glycine,

proline, and arginine, AMC = 7-amino-4-methylcoumarin), Z-FR-AFC (FR=

phenylalanine and arginine, AFC= 7-Amino-4-trifluoromethylcoumarin), and Z-VVR-

AMC (VVR= valine, valine, and arginine), and three small chromogenic peptide

substrates, Suc-AAPF-pNA (N-succinyl-alanine-alanine-proline-phenyalanine-p-

nitroanilide), H-Arg-pNA.2HCl, and Z-FR-pNA.HCl (all obtained from Enzo® Life

Sciences, Farmingdale, NY) using a fluorometer (Wallac 1420, Victor3TM- PerkinElmer).

However, Z-FR-AFC was found to be the best substrate and was used for all

subsequent assays. The 96-well plate format (Corning® - Corning, NY - 96 Well Flat

Clear Bottom Black Polystyrene TC-Treated Microplates, to prevent well-to-well

crosstalk) was used contained a final assay volume of 150 µl per well. The fluorescent

assays were conducted at excitation/emission wavelengths of 355 nm/535 nm, 25 °C,

with a final substrate concentration of 12.5 µM, whereas the chromogenic assays were

conducted at an absorbance of 405 nm with final substrate concentration of 80 µM.

Fluorescence was monitored at 5 min intervals for up to 120 min. To find the optimum

pH, the following buffers were used at the variously stated pH range: 50 mM sodium

acetate buffer (pH 3-4), 50 mM MES buffer (pH 5-6), and 50 mM Tris-HCl (pH 7-9). For

proteinase inhibitor assays, the enzyme-inhibitor mixtures (without substrate) were pre-

incubated at room temperature for at least 10 min prior to the assay. The assays were

performed in duplicates/triplicates and repeated1-3 times. The inhibition of DaCatL1

was assessed using the protease inhibitors within their physiological working

concentrations: E-64 (L-trans-Epoxysuccinyl-leucyamido (4-guanidino) butane), Z-

48

FY(tBu)DMK (N-benzyloxycarbonyl-phenylalanyl-t-butyl-tryrosyl diazomethylketone),

CA074, cystatin and chymostatin at 50 µM each and aprotinin at 50 nM.

D. abbreviatus gut in vitro assays: To find the cysteine protease(s) activity of

D. abbreviatus cathepsins, crude cathepsins from the guts of 5 actively feeding 6-week-

old larvae were obtained by homogenizing surgically removed midguts in deionized

water (1 larval gut per 100 µl of water). Debris was removed by centrifugation (14,000

rpm at 4 °C) and supernatants saved as 25 µl aliquots. Enzymatic assays were

conducted as described with Z-FR-AFC (12.5 µM) in the presence of cathepsin B

specific inhibitor CA074 (50 µM). Single gut aliquots were removed as needed and

diluted 1:100 for enzymatic assays. One microliter was used per well in a total assay

volume of 150 µl.

To investigate the role TritonTM X-100 might have on protein folding and therefore

the activity of cathepsins and the interaction with the peptide inhibitors, an experiment

was performed in which the guts of four six-week-old larvae were extracted into buffer

containing TritonTM X-100 (1%) and another without TritonTM X-100. Enzymatic assays

were performed at acidic (pH 5.5), neutral (pH 7), and basic (pH 8) conditions using the

substrate Z-FR-AFC, with the gut extracts including either of the two inhibitors. The

cathepsin B inhibitor (CA074) was included at 50 µM. To eliminate the possibility that

TritonTM X-100 affected the interaction with the inhibitors, I tried to remove TritonTM X-

100 using HiPPR® Detergent Removal Spin Columns (Thermo Scientific).

Results

Modeling of Cathepsin L1 and CPI1

DaCatL1 consists of a short N-terminal propeptide with three alpha-helices,

linked to a longer C-terminal catalytic domain which possesses two catalytically active

49

residues Cys120 and His255 located at the centre of the catalytic groove (Figure 2-1 A).

Both domains are organized in such a way that the propeptide completely blocks the

catalytic groove of the active domain, inhibiting the proteinase activity of cathepsin L1

(Figure 2-1 B, C). After cleaving the propeptide from the catalytic groove, DaCatL1

recovers its proteolytic activity.

CPI1 comprises three alpha-helices similarly organized as in the propeptide of

DaCatL1. The amino acid sequences of the propeptide and CPI1 share high similarity (~

70%) but moderate percentages of identity (~25%). According to these common

structural organizations, the modelled CPI1 exhibits similar overall three-dimensional

conformation to the natural propeptide of DaCatL1 with a long C-terminal tail at the end

of the alpha-helical core structure (Figure 2-1 D). The C-terminal tail allows CPI1 to bind

in the catalytic groove of DaCatL1 (Figure 2-1 E) and to block its enzymatic activity. The

enzymatic inhibition is due to steric hindrance at the catalytic groove that prevents the

substrate from interacting with the active residues Cys120 and His255 at the active

groove (Figure 2-1 F). CPI1 interacts with the activated DaCatL1 through a network of

hydrogen bonds forming between the core structure of CPI1 and the upper part of

DaCatL1. The endogenous CPI1 offers D. abbreviatus an efficient mechanism to

regulate cathepsin L activity in the gut by mimicking the natural propeptide which

maintains DaCatL1 inactive.

Synthesis and Purification of DaCatL1

Although the E. coli cloning, synthesis, and subsequent purification of CPI1 and

its mutant (CPI1mut) were straight-forward with standard techniques, the production of

DaCatL1 was very challenging. Previously, unsuccessful attempts at synthesizing D.

abbreviatus cathepsin L2 were made using the yeast expression system of Pichia

50

pastoris (pPICz-alpha, Invitrogen), a Human In Vitro Translation Kit (Thermo Scientific),

and DaCatL1 in E. coli (pET200 Directional TOPO® Expression Kit, Life Technologies) .

To overcome these difficulties, DaCatL1 was expressed in an E. coli system fused to a

GST tag. Syntheses of recombinant proteins were induced with IPTG in DaCatL1-

transformed E. coli cells, alongside non-transformed control cells. Crude bacterial

extracts were obtained and run on SDS-PAGE (Figure 2-2, lanes a-b). DaCatL1 fusion

product was affinity purified (Figure 2-2, lane c) followed by size-exclusion

chromatography of the mature DaCatL1 at pH 8 (Figure 2-2, lane d). In the SDS-PAGE

analysis of DaCatL1 (Figure 2-2), the inactive fusion DaCatL1 with GST tag is present

at the molecular weight size of 60KDa and the active DaCatL1 (Mr 23 KDa) is also

present. Monoclonal antibodies against His6-tag proteins identified DaCatL1 from

transformed E. coli cells but not the non-transformed control (Figure 2-3).

Although the GST tag enabled high synthesis of DaCatL1, the recombinant

enzyme was unstable and readily precipitated out of solution, regardless of the different

conditions that were employed, such as low salt LB cultures, induction and expression

at various temperatures (25 – 37 °C), and the incorporation of the surfactants

(amphipol, CHAPS, OGNG and TritonTM X-100 at different concentrations) during lyses.

DaCatL1 only remained soluble in TritonTM X-100 (1%).

The mass spectrometry analysis showed 34 unique peptides (yellow in Figure 2-

4) indicating 100% probability that the expressed purified protein is DaCatL1 fused to

GST and His6 (end of sequence).

Activated DaCatL1 was eluted under the second protein peak (red arrow in

Figure 2-5) using size-exclusion chromatography and Tris-HCl (pH 8) buffer. The eluted

51

protein Mr was found to be 23 kDa by SDS-PAGE and the protein showed Cathepsin L

activity when incubated with its substrate ( Z-FR-AFC) at pH 8 (Figure 2-2, lane d). The

first protein peak (blue arrow in Figure 2-5) is the non-activated moiety of Cathepsin L

(Mr 60 kDa) after analysis by SDS-PAGE (results not shown).

Synthesis and Purification of CPI

Two forms of CPI 1 were expressed and purified. The first form was the wild

type CPI1 that was discovered in D. abbreviatus. The second form (CPI1mut) was a

mutated CPI1 in which Lys101 was replaced by Cys to form an S-S bond between

Cathepsin L Cys120 and CPI1mut Cys101. The inhibitor was extracted from bacterial

lysates, purified by affinity chromatography on Ni-NTA resin, and then purified to

apparent homogeneity by size exclusion chromatography. At each purification step,

CPI1mut was assayed for homogeneity by SDS-PAGE (Figure 2-6a-d). After size

exclusion chromatography, the inhibitor exhibited apparent homogeneity and was also

assayed by monoclonal antibodies to the His6-tag, proving that staining the purified His6-

CPI1mut with coomassie brilliant blue or analyzing with His6 antibodies detect the same

protein of Mr 11 kDa (Figure 2-6 d and e).

Cathepsin L1 Substrate Specificity and pH Optimum

DaCatL1 hydrolyzed the general cathepsin L, B and H substrate, Z-FR-AFC

(Figure 2-7). Thus, all activity assays were carried out in the presence of 50 µM CA074

(cathepsin B inhibitor). The activity assay showed that DaCatL1 is more active at pH 8

than at pH 5.5. DaCatL1 showed optimal activity at pH 8 as determined with fluorogenic

substrate Z-FR-AFC at 25 ºC (Figure 2-8). The fluorogenic substrate is more sensitive

in the bioassay than its chromogenic counterpart, Z-FR-pNA (results not shown) or Z-

GPR-AMC and Z-VVR-AMC (Figure 2-7). These results indicate that DaCatL1 is an

52

alkaline cathepsin in contrast to the majority of cathepsins that show optimal activity at

acidic pH range. The cathepsin G substrate SucAAPF-pNA (also cleaved by

chymotrypsin-like proteases) and HR-pNA (Cathepsin H and other aminopeptidases)

were not cleaved by DaCatL1 (data not shown).

Effect of Inhibitors on Cathepsin L1 Activity

The activity of DaCatL1 was assessed at pH 8 (optimal enzyme activity, Figure 2-

8) in the absence or presence of known protease inhibitors by incubating the enzyme

and the inhibitors together for at least 10 min and conducting the assay for 120 min.

Aprotinin (serine protease inhibitor) at 50 nM weakly inhibited DaCatL1; it inhibited

DaCatL1 activity almost completely at a 1000-fold higher concentration (data not

shown). In the presence of the general cathepsin inhibitor (excluding cathepsin C and

H) E-64 (50 µM), the activity was inhibited by 26.6%, whereas the specific cathepsin L

inhibitor Z-FY(tBu)DMK (50 µM) caused higher inhibition (40.6%). The specific

cathepsin B inhibitor (CA074) at 50 µM inhibited the enzyme by 25.3%. The general

protease inhibitors cystatin and chymostatin (50 µM each), which also inhibit cathepsin

L, were more effective, inhibiting DaCatL1 activity, by 47.5% and 85.6%, respectively

(Figure 2-9).

To determine if DaCatL1 is affected by CPI1 and CPI1mut, the enzyme was

separately incubated with each inhibitor and its activity followed in the presence and

absence of the inhibitors (Figure 2-10). As inhibitor concentration increased from 1

µMto 10 µM, an increase in inhibition was recorded. Both inhibitors were equally

effective. However, the inhibition of the enzyme by both inhibitors did not reach 50%

53

even in the presence of 10 µMof CPI1 or CPI1mut, indicating that DaCatL1 is only

partially inhibited by CPI1 inhibitors at pH 8.0.

Sequence Alignment of Cathepsin L1 and Human Cathepsins

Alignment of amino acid sequences of DaCatL1 and human cathepsins showed

similarities among the different cathepsin sequences (Figure 2-11). Active site residues

of DaCatL1 and human Cathepsin L1 (HumCTL1): C, H, Q and N, are conserved in all

the different cathepsins. Therefore, folding of the protein molecules is important to

achieve proper catalytic activity of these proteases. Thus, substrate specificities and

substrate binding sites (Turk et al., 2000) are important in distinguishing the different

cathepsins even though high conservation at the catalytic amino acids is observed

(Figure 2-11).

D. abbreviatus Gut in vitro Assays

In acidic and neutral conditions, TritonTM X-100 negatively affected the catalytic

activity of the gut cathepsin Ls, whereas the inhibition of the enzymatic activity of the gut

cathepsin Ls with either inhibitor is not affected by TritonTM X-100 (Figure 2-12). The

enzyme-inhibitor interaction is predominantly influenced by the pH of the incubation

mixture. Since the catalytic activity of the enzyme itself is affected by TritonTM X-100, it

is not clear if the surfactant has an effect on the binding of CPIs to the enzyme. At the

basic pH, TritonTM X-100 does not have an effect on the enzymatic activity of the basic

Cathepsin L or the binding of CPI to it.

Removing TritonTM X-100 using HiPPR® Detergent Removal Spin Columns

(Thermo Scientific) had no effect on the enzymatic activity observed earlier (data not

shown). However, protein recovery of the enzyme after the spin column step was low.

This may be because of the instability of DaCatL1 in the absence of TritonTM X-100.

54

Both CPI inhibitors were very effective inhibitors of digestive cysteine proteases

at pH 4-7 as they inhibited 80- 90% of the gut cathepsin L activities (Figure 2-13);

however, at pH 8 the inhibitors inhibited13- 25% of the gut enzymatic activity. This

suggests that DaCatL1 is a basic cathepsin L in the gut of D. abbreviatus.

Discussion

The interaction between cathepsin L 1 and CPI1 were modeled by Pierre Rougé

(Université de Toulouse, Institut de Recherche pour le Développement). Using Rouge’s

3D models, CPI1 was modified by mutating one of the amino acids that is in close

proximity to Cathepsin L unbridged cysteine in order to bind CPI to Cathepsin L using S-

S bonding. This is postulated to make the binding between CPI and Cathepsin L

irreversible in the hope of using this strategy in transgenic citrus. Cathepsin L protease

inhibitors expressed in the roots of citrus will prevent food digestion in the gut of larval

D. abbreviatus.

The working hypothesis was formed from gut transcriptome studies that DaCatL1

is a major digestive enzyme (R. Shatters, unpublished data). Furthermore, based on the

predominance of high activity of cathepsins at acidic pH, DaCatL1 was postulated to be

active at acidic pH, where the pro-segment of the molecule is cleaved. In the posterior

midgut (alkaline pH), DaCatL1 molecules are bound by modulators of their activity,

cystatins – cysteine protease inhibitors (Otto and Schirmeister, 1997), to protect them

from degradation and therefore conserve the insects energies in resynthesizing new

DaCatL1 molecules. Modeling the interaction between DaCatL1 and CPI (Figure 2-1)

indicates a reversible interaction via hydrogen bonding between specific amino acid

residues on the surface CPI and in the active-site pocket of DaCatL1 which is pH

dependent.

55

The sequence of events in the gut of D. abbreviatus was therefore suggested as

follows. As food particles move through the insect midgut, passing from the anterior end

to the posterior end, molecules (including enzymes) move back and forth, across the

peritrophic membrane to improve the energy efficiency of digestion. Cathepsin L1,

postulated to be active in the anterior region, is therefore available to cleave peptide

bonds. However, in the posterior midgut, DaCatL1 is protected from degradation by

CPI1, by sitting in the active site pocket of the enzyme. A proportion of the DaCatL1-

CPI1 complex is found again in the anterior midgut as it cycles through the gut, where

hydrogen bonds, loosely holding the enzyme and the inhibitor together, are broken to

release active DaCatL1 molecules. This is a measure by the insect to conserve its

energy from the need to resynthesize new enzyme.

To verify this hypothesis, it was important to synthesize and purify recombinant

DaCatL1 and inhibitors to characterize their kinetics. Going a step further, a mutant form

of the inhibitor was synthesized by modifying the amino acid residue directly involved in

the interaction with DaCatL1, such that the chemical bond between this mutant peptide

and DaCatL1 is changed from the weak hydrogen bond to a stronger covalent bond that

is not pH-dependent. Success in this approach was postulated to make available a

more potent inhibitor of cathepsin and a better candidate toxin in a transgenic control

strategy.

The copious synthesis of DaCatL1 in an E. coli system was accomplished by

synthesizing it as a fusion product with a GST tag. However, DaCatL1 was very

unstable in solution; the cause of this instability is not clear. Perhaps differences in

protein processing in a prokaryotic system may contribute to this instability. Still, other

56

polypeptide moieties could be employed that may keep DaCatL1 soluble and retain its

activity, such as the maltose binding protein (MBP) and the small ubiquitin-like modifier

(SUMO) (Kapust and Waugh, 1999; Butt et al., 2005). Of the surfactants employed to

improve the solubility of DaCatL1, only TritonTM X-100 (1%) was effective, though it is

not clear if it affected the characterization of DaCatL1. The list of surfactants assayed is

by no means exhaustive, and other options may be more useful. Additionally, numerous

alternative recombinant protein synthesis technologies are available (Hunt, 2005), but

the expression of heterologous proteins in the E. coli system is simple and straight-

forward.

The characterization of DaCatL1 presents confounding data. DaCatL1

preferentially cleaved Z-FR-AFC (Figure 2-7), which has a hydrophobic phenylalanine at

P2. However, DaCatL1 had optimum activity at pH 8 (Figure 2-8). This is unusual of the

cathepsins in the MEROPS database (Rawlings et al., 2012). Reports of basic cysteine

proteases are rare (Otto and Schirmeister, 1997); most have been found to have

optimum activity in the acidic pH range (pH 5-6.5). Though Z-FR-AFC is an established

cathepsin L substrate (Tchoupé et al., 1991), cathepsins B and H have affinities also for

it. Specific small-molecule inhibitors also aid the characterization of cathepsins. The

general cathepsin inhibitor E-64 inhibits the activities of almost all cysteine proteases,

including cathepsins B, H, and L (Barrett et al., 1982; Turk et al., 2012), but not

cathepsin C (Rozman-Pungercar et al., 2003). Meanwhile, the inhibitor CA074 and its

analogs specifically inhibit cathepsin B (Bogyo et al., 2000), whereas Z-Phe-Tyr(tBu)-

diazomethylketone and analogs specifically inhibits cathepsin L (Shaw et al., 1993).

Some of these inhibitors, among others, were used to characterize DaCatL1. To rule out

57

cathepsin B, the specific cathepsin inhibitor, CA074, was assayed against DaCatL1. In

the presence of CA074 (50 µM), DaCatL1 still retained 75% of its activity (Figure 2-9).

However, the general cathepsin inhibitor E-64 and the specific cathepsin L inhibitor Z-

FY-(tBu)DMK only showed limited inhibition of DaCatL1 (Figure 2-9).The general

cathepsin inhibitor chymostatin was effective at inhibiting DaCatL1 (Figure 2-9),

perhaps because it also inhibits serine proteases that are active at a basic pH.

Additionally, aprotinin, a sole serine protease inhibitor, did not inhibit DaCatL1 at the

same concentration range at which it effective against serine proteases. It is worthy to

re-emphasize here that none of the substrates: Z-GPR-AMC, Z-VVR-AMC, Suc-AAPF-

pNA and H-Arg-pNA.2HCl were cleaved by DaCatL1. Crucially, mass spectrometry

analysis (Figure 2-4) of SDS-PAGE bands cut out for analysis (Figure 2-2), put the

synthesis and purification (Figure 2-5) of DaCatL1 beyond any doubt. DaCatL1, as

synthesized by an E. coli strain, is best described as a cryptic cysteine protease.

These data beget the question: Does DaCatL1 have chymotrypsin-like activity?

When the chymotrypsin-specific inhibitor TPCK (tosyl phenylalanyl chloromethyl ketone)

was incubated with DaCatL1 in the presence of Z-FR-AFC, DaCatL1 was not inhibited;

rather it exhibited full cysteine protease activity (data not shown). This was also the

case with a positive control using a commercially recombinant human cathepsin L1. It is

therefore safe to conclude that DaCatL1 does not have chymotrypsin-like activity.

Aligning DaCatL1 with other sequenced cathepsins revealed the same active site

residues: C, H, Q and N (Figure 2-11), yet these cathepsins have different substrate

preferences (Rawlings et al., 2012). Proper protein folding is therefore a prerequisite for

the activity/function of these proteases to determine their kinetic characteristics. Access

58

of substrates to substrate binding sites is important in distinguishing between the

cathepsins (Turk et al., 2000).TritonTM X-100 was the only surfactant that kept the

protein stable and active in solution. Therefore proof that the enzyme is properly folded

in its native form has to wait for future X-ray crystallographic studies.

Cathepsin L protease inhibitor wild type (CPI1) and its mutant (CPI1mut) were

successfully purified and effectively inhibited the activity of gut cathepsins from D.

abbreviatus larvae at acidic to neutral pH. Even though some inhibition of gut

cathepsins was observed at basic pH with either CPI, it was low (25% or less). This

was also true with DaCatL1 and it is likely that the basic cathepsins are uniquely

regulated and may have their own specific modulators/regulators. The inhibitory potency

of Cathepsin L propeptides were found to be pH dependent and were highly potent at

acidic pH (Carmona et al., 1996).

I was unable to show that CPI1mut is a better candidate for inhibition of D.

abbreviatus acidic cathepsins than the wild-type. The cysteine residues (for instance,

active site cysteine of DaCatL1 and the mutated cysteine residue on CPI1mut) are

probably too far apart to establish an effective S-S bond. X-ray crystallographic data

may allow a better prediction of an amino acid to mutate on the C-terminus of CPI1 to

get it close to the active site of D. abbreviatus digestive cathepsins. Possibly, an

opportunity of evaluating a better mutant was missed due the failure to synthesize

active D. abbreviatus cathepsin L2 in E. coli.

A re-evaluation of the process of digestion in D. abbreviatus may also be

necessary. It is apparent, however, that the DaCatL1-CPI interaction obtained in this

study is not the one modeled and hypothesized. DaCatL1 is entirely different from the

59

acidic cathepsins that are modulated by CPI1. This notwithstanding, the strategy of

mutating endogenous regulators of enzymatic activity to make for better inhibitors is a

feasible and attractive possibility to protect crops. Cathepsin L protease inhibitors are a

potential alternative in a transgenic strategy for crop protection in citrus rootstock

against D. abbreviatus as a component of an integrated IPM system.

60

Figure 2-1. Modelling of the interaction between Diaprepes abbreviatus cathepsin L1

(DaCatL1) and cathepsin L protease inhibitor 1 (CPI1). A shows the overall three-dimensional conformation of DaCatL1 and CPI1. DaCatL1 propeptide (PP) and the catalytic domain are colored violet red and yellow, respectively and the catalytic residue of DaCatL1 C120 and H255 red and blue, respectively. B shows the catalytic groove (black dashed lines) of the activated DaCatL1 and its propeptide PP. C is an enlarged view of the catalytic residues C120 and H255 at the C-terminal tail of the propeptide (PP) bound in the catalytic groove of DaCatL1. D is a ribbon diagram showing the docking of cathepsin inhibitor CPI1 from D. abbreviatus to the activated cathepsin DaCatL1. The inhibitor and the activated cathepsin L1 are colored violet red and yellow, respectively. E shows the molecular surfaces and the binding of CPI 1 into the catalytic groove (black dashed lines) of the activated DaCatL1. The catalytically active C120 and H255 residues are also shown. F is an enlarged view of the catalytic residues C120 and H255 and the C-terminal tail of the cathepsin inhibitor CPI11 binding to the catalytic groove of the activated DaCatL1.

A B

C

D E

F

CPI CPI

CPI

PP PP

PP C120

C120

C120

C120 C120

C120

H255 H255

H255

H255 H255

H255

61

Figure 2-2. SDS-PAGE analysis of recombinant cathepsin L1 purification. Lane on far left has molecular weight markers. Lanes a= control, b= crude bacterial extract, c= affinity (Ni-NTA) purified enzyme, d= activated affinity purified enzyme after size exclusion chromatography and concentration. The red box highlights the recombinant fusion protein, Mr 60KDa before activation and the red arrows shows an activated DaCatL1.

250

75

50

25

20

KDa a b c d

62

Figure 2-3. Western blot analysis of His- tagged DaCatL1. Left lane is a control of

protein extract from E.coli cells with a plasmid not carrying DaCatL1. Right lane protein extract from DaCatL1 in transformed E. coli cells showing two expected bands of Mr 60KDa (GSTHis6-DaCatL1) and Mr 23 KDa (activated DaCatL1 without the GST tag). Bacterial cells were induced with IPTG (as above).

Figure 2-4. Mass spectrometry analysis of DaCatL1.

Control DaCatL1

60 KDa

23 KDa

63

Figure 2-5. Size-exclusion purification of DaCatL1. Activated DaCatL1 was eluted in the second peak (red arrow) during size-exclusion chromatography with Tris-HCl (pH 8). The first peak (blue arrow) was resolved on SDS-PAGE as the 60 KDa fusion product indicated in Figure 2-2.

Figure 2-6. SDS PAGE (lanes a-d) and western blot analysis (lane e) of cathepsin L protease inhibitor1 mutant (CPI1mut). Lane a: control lane of crude bacterial extract; Lane b: CPI1mut transformed E.coli crude protein extract showing 11KDa bands in the red box (the calculated size of the His6- tagged protein); Lane c: affinity (Ni-NTA) purified protein; Lane d: size exclusion chromatography purification after affinity chromatograph (lane c); Lane e: Monoclonal antibody identification of CPI1mut band on nitrocellulose membrane was visualized by fluorescence.

25

10 11KDa

a b c d e

64

Figure 2-7. Fluorogenic substrate preference of DaCatL1. Activity is given as RFU = relative fluorescent unit.

Figure 2-8. pH activity profile of recombinant DaCatL1 using Z-FR-AFC

65

Figure 2-9. Inhibition of DaCatL1 by small-molecule protease inhibitors.

Figure 2-10. Inhibition of cathepsin L protease inhibitors on the activity of DaCatL1. Enzyme activity was followed at pH 8, 25 °C using 2.25 µg of synthesized recombinant enzyme per assay.

66

DaCatL1 ----MKVFIAACLLVAVSATVLEETGVK------FQAFKLKHGKTYKNQVEETARFNIFK

HumCTH MWATLPLLCAGAWLLGVPVCGAAELCVNSLEKFHFKSWMSKHRKTYS-TEEYHHRLQTFA

HumCTK ------MWGLKVLLLPVVS---FALYPEEILDTHWELWKKTHRKQYNNKVDEISRRLIWE

HumCTS ------MKRLVCVLLVCSSAV-AQLHKDPTLDHHWHLWKKTYGKQYKEKNEEAVRRLIWE

HumCTL1 ------MNPTLILAAFCLGIASATLTFDHSLEAQWTKWKAMHNRLYG-MNEEGWRRAVWE

HumCTB ---------------------------------MWQLWASLCCLLVLANARSRPSFHPLS

: :

DaCatL1 DNLR-AIEQHNVLYEQGLVSYKKGINRFTDMTQEEFRAFLTLSSSKKPHFN---TTEHVL

HumCTH SNWR-KINAHN----NGNHTFKMALNQFSDMSFAEIKHKYLWSEPQNCSAT---KSNYLR

HumCTK KNLK-YISIHNLEASLGVHTYELAMNHLGDMTSEEVVQKMTGLKVPLSHSRSNDTLYIPE

HumCTS KNLK-FVMLHNLEHSMGMHSYDLGMNHLGDMTSEEVMSLMSSLRVPSQWQR--NITYKSN

HumCTL1 KNMK-MIELHNQEYREGKHSFTMAMNAFGDMTSEEFRQVMNGFQNRKPRKG---KVFQEP

HumCTB DELVNYVNKRNTTWQAGHNFYNVDMSYLKRLCG----TFLGGPKPPQRVMF----TEDLK

.: : :* * : :. : :

DaCatL1 TGLAVPDSIDWRTKG-QVTGVKDQGNCGSCWAFSVTGSTEAAYYRKAGKLVSLSE--QQL

HumCTH GTGPYPPSVDWRKKGNFVSPVKNQGACGSCWTFSTTGALESAIAIATGKMLSLAE--QQL

HumCTK WEGRAPDSVDYRKKG-YVTPVKNQGQCGSCWAFSSVGALEGQLKKKTGKLLNLSP--QNL

HumCTS PNRILPDSVDWREKG-CVTEVKYQGSCGACWAFSAVGALEAQLKLKTGKLVSLSA--QNL

HumCTL1 LFYEAPRSVDWREKG-YVTPVKNQGQCGSCWAFSATGALEGQMFRKTGRLISLSE--QNL

HumCTB LPASFDAREQWPQCP-TIKEIRDQGSCGSCWAFGAVEAISDRICIHTNAHVSVEVSAEDL

:: :. :: ** **:**:*. . : . :. :.: ::*

DaCatL1 VDCSTDIN--AGCNGGYLDETFTYVKS-KGLEA---------------------------

HumCTH VDCAQDFN-NHGCQGGLPSQAFEYILYNKGIMG---------------------------

HumCTK VDCVSE---NDGCGGGYMTNAFQYVQKNRGIDS---------------------------

HumCTS VDCSTEKYGNKGCNGGFMTTAFQYIIDNKGIDS---------------------------

HumCTL1 VDCSGPQ-GNEGCNGGLMDYAFQYVQDNGGLDS---------------------------

HumCTB LTCCGSMC-GDGCNGGYPAEAWNFWTRKGLVSGGLYESHVGCRPYSIPPCEHHVNGSRPP

: * ** ** :: : : .

DaCatL1 ---ESTYPYKGTDGSCKYSASKVVTKVSGHKSLKSE-DENALLDAVGNVGPVSVAIDATY

HumCTH ---EDTYPYQGKDGYCKFQPGKAIGFVKDVANITIY-DEEAMVEAVALYNPVSFAFEVTQ

HumCTK ---EDAYPYVGQEESCMYNPTGKAAKCRGYREIPEG-NEKALKRAVARVGPVSVAIDASL

HumCTS ---DASYPYKAMDLKCQYDSKYRAATCSKYTELPYG-REDVLKEAVANKGPVSVGVDARH

HumCTL1 ---EESYPYEATEESCKYNPKYSVANDTGFVDIPK--QEKALMKAVATVGPISVAIDAGH

HumCTB CTGEGDTPKCSKICEPGYSPTYKQDKHYGYNSYSVSNSEKDIMAEIYKNGPVEGAFSVYS

: * . :.. . *. : : .*:. ....

DaCatL1 --LSSYESGIYEDDWCS--PSELNHGVLVVGYGTS----NGKKYWIVKNSWGGSFGESGY

HumCTH D-FMMYRTGIYSSTSCHKTPDKVNHAVLAVGYGEK----NGIPYWIVKNSWGPQWGMNGY

HumCTK TSFQFYSKGVYYDESCNS--DNLNHAVLAVGYGIQ----KGNKHWIIKNSWGENWGNKGY

HumCTS PSFFLYRSGVYYEPSCT---QNVNHGVLVVGYGDL----NGKEYWLVKNSWGHNFGEEGY

HumCTL1 ESFLFYKEGIYFEPDCSS--EDMDHGVLVVGYGFESTESDNNKYWLVKNSWGEEWGMGGY

HumCTB D-FLLYKSGVYQHVTGEM---MGGHAIRILGWGVE----NGTPYWLVANSWNTDWGDNGF

: * *:* .*.: :*:* .. :*:: ***. .:* *:

DaCatL1 FRLLRGK-NECGVAEDTVYPII--------- 322

HumCTH FLIERGK-NMCGLAACASYPIPLV------- 335

HumCTK ILMARNKNNACGIANLASFPKM--------- 329

HumCTS IRMARNKGNHCGIASFPSYPEI--------- 331

HumCTL1 VKMAKDRRNHCGIASAASYPTV--------- 333

HumCTB FKILRGQ-DHCGIESEVVAGIPRTDQYWEKI 339

. : :.: : **:

Figure 2-11. Alignment of DaCatL1 with human cathepsin sequences. Highlighted in yellow are the active site residues.

67

Figure 2-12. Influence of TritonTM X-100 (1%) on the interaction between D. abbreviatus gut cathepsins and the cathepsin L protease inhibitors CPI1 and CPI1mut.

68

Figure 2-13. Inhibition of D. abbreviatus larval cathepsins by cathepsin L protease inhibitors

69

CHAPTER 3 A METHOD FOR CONTROLLING DIAPREPES ABBREVIATUS LARVAL FEEDING

ON CITRUS USING A BACILLUS THURINGIENSIS CYTOLYTIC TOXIN

Diaprepes abbreviatus (Linnaeus) rapidly established itself as an important pest

of citrus in USA after its accidental introduction from Puerto Rico in the 1960s (Lapointe

et al., 2007). There have been several attempts to manage this pest in citrus groves, but

they have failed for numerous reasons. A natural insect control strategy that has been

widely exploited for insect control in agriculture is based on the production of insecticidal

proteins, known as delta endotoxins, located in parasporal inclusion bodies of the soil

bacterium, Bacillus thuringiensis (Lacey et al., 2001; Crickmore, 2006; Bravo et al.,

2011). The delta-endotoxins are put into two groups: Crystal (Cry) and Cytolitic (Cyt),

and act differently but culminate in the disruption of gut integrity. The Cry toxins bind

specific receptors and results in pores in the gut epithelium, whereas the Cyt toxins act

in a detergent-like manner to destroy the integrity of the gut epithelium (De Maagd et al.,

2001; Manceva et al., 2005). Numerous toxins from several B. thuringiensis strains

show specific activity to members of Diptera, Lepidoptera, and Coleoptera. Whereas the

Cyt toxins predominantly show dipteran toxicity (Bravo et al., 2007), a few Cyt toxins

(Cyt1Aa and Cyt2Ca) have been found to have toxicity to some coleopteran species

(Soberón et al., 2013), including Cyt2Ca1, which is active against D. abbreviatus

(Weathersbee et al., 2006). Formulations of Bt are used effectively as spray

applications in horticulture, but limitations such as rapid environmental degradation and

the subterranean feeding habit of D. abbreviatus larvae, means such a tactic will be

impractical in citriculture.

Since 1996, significant advancement of the use of Bt as bio-control agents

(Crickmore, 2006) has been the commercial deployment insect-resistant transgenic

70

crops expressing a variety of Cry toxins. However, this has yet to be introduced in

commercial citrus fruit trees. The transgenic crop varieties have been rapidly adopted

by farmers providing insurmountable proof of their resilience against destruction from

feeding damage by pests. Because of the way citrus is commercially cultivated, by

grafting fruiting scions onto resistant rootstock cultivars, it lends itself to a transgenic

citrus rootstock strategy to protect against D. abbreviatus larvae. Citrus rootstocks can

thus be engineered for resistance to D. abbreviatus, but not the scions, leaving fruits

without transgenic genes. Evidence that recombinant products are not expressed in

fruits will boost consumer acceptance. Citriculture can therefore benefit from the

transgenic technology. Thus, the citrus variety ‘Carizzo’ (a hybrid of Citrus sinensis × P.

trifoliata), was engineered to express a recombinant gene product, Cyt2Ca1.

This chapter presents results of work done with transgenic Cyt2Ca1 citrus to

identify transformational events carrying and expressing the recombinant gene product.

A transformational event is a genetically engineered organism (citrus in this case) and

all identical clones resulting from a transformation. In addition, transgenic Cyt2Ca1

citrus was challenged with aggressively feeding D. abbreviatus larvae to determine the

plant’s resistance to feeding damage.

Materials and Methods

Citrus Plants

Several batches of transgenic citrus plant cohorts engineered with Cyt2Ca1 were

available from USDA-USHRL. A total of 75 transgenic citrus cohorts (scions were

transgenic and grafted to non-engineered rootstocks), making up batch 1, were

screened by qPCR on leaf specimens for presence of the Cyt2Ca1 insert. Cuttings of

qPCR validated parental transgenic citrus (Batch 2) were made from Batch 1 and used

71

for transcription analysis by q-RT-PCR. Furthermore, cuttings (Batch 3) were made from

selected batch-2 transgenic citrus plants and used for potted-plant bioassays.

Primers

Primers listed in Table 3-1 were designed and purchased from IDT® (San Diego,

CA) and used for q-PCR and q-RT-PCR amplifications. Primers BtCyt-f and BtCyt-r

were used to amplify Cyt2Ca1, CitCD-f and CitCD-r were used to amplify citrus

dehydrogenase (qPCR reference genes), whereas CitUbi-f and CitUbi-r, CitAct-f and

CitAct-r, and Cit18S-f and Cit18S-r were used to amplify ubiquitin (HarvEST:

UCRCP01), actin (HarvEST: C0LQ97), and 18S RNAs (GenBank: AF206997.1),

respectively (q-RT-PCR reference genes).

Table 3-1. Primers used for qPCR and q-RT-PCR

Name Sequence (5’- 3’) Tm (melting temperature)

°C

BtCyt-f ATG TTC TTC AAC CGC GTT ATC 54.0 BtCyt-r TTA GCT GTT GCT GCA GAT TTT 54.9 CitCD-f TGA GTA CGA GCC GAG TGT TG 56.7 CitCD-r CTG GTG GAT CGG TGA AGT TT 55.2 CitUbi-f TCG CCG ATT ACA ACA TTC AA 52.4 CitUbi-r AGA GGA AAT TAG GCC CAA GC 54.9 CitAct-f AAA TCA AAC CCC AGG CTT CT 54.6 CitAct-r ACC CTT GCG TCG TAC AGT TC 57.4 Cit18s-f AAA CGG CTA CCA CAT CCA AG 55.0 Cit18s-r CCT CCA ATG GAT CCT CGT TA 53.2

Nucleic Acid Extractions

DNA Extraction: Two leaf samples were randomly obtained from batch-1

transgenic citrus plants, and cleaned with RNase™ AWAY (Thermo Scientific, Waltham,

MA). Leaves were thoroughly rinsed with de-ionized water, and excess water was

blotted off. Leaf midribs were excised and finely chopped (a new blade was used per

sample to prevent DNA cross-contamination) for genomic DNA extractions using the

72

FastDNA® Green SPIN Kit (mpBIO, Solon, OH) according to the manufacturer’s

protocol. The purity and concentration of DNA was quantified by NanoDrop™ 1000

(Thermo Scientific).

RNA Extraction: Two leaf samples of qPCR-validated batch-2 Cyt2Ca1 citrus

were obtained and processed as described previously but the midrib-excision step was

skipped. Samples were placed in zipper-sealed sample bags and flash frozen with liquid

nitrogen. The samples were stored at -80 °C until they were ready to be used for RNA

extractions. The TRI Reagent® (Sigma-Aldrich, St Louis, MO) extraction protocol (with

modifications) was used for RNA extractions from leaf samples taken from F1 cohorts.

All materials were ensured to be RNase-free with RNAse AWAY (Thermo Scientific).

The leaf samples were ground to a fine powder under liquid nitrogen with a

mortar and pestle. With a spatula, 150 mg of pulverized leaves were placed into 2-mL

screw-cap tubes. One mL of TRI Reagent® was added to the tubes and the mixture

vortexed then, and incubated at room temperature for 5 min. After incubation, 0.2 mL of

chloroform was added to the mixture, and the mixture was vortexed and incubated for

10 min at room temperature. Proteins and DNA were precipitated and pelleted by

centrifugation at high speed (12,000 rpm for 10 min at 4 °C in a refrigerated bench-top

centrifuge- Eppendorf 5702 R). The partitioned aqueous phase was transferred into a

fresh 1.5-mL tube (DNA- and RNA-free). Five hundred µl of isopropanol were added per

1 mL TRI reagent, vortexed briefly, and incubated for 10 min. The solution was

centrifuged at high speed, and the supernatant was discarded. To wash the pellet, 1 mL

of cold 75% ethanol was added to the pellet and the tube inverted gently. The ethanol

was removed after centrifugation at high speed for 5 min and RNA pellet air-dried for

73

approximately 10 min. The RNA pellet was then re-suspended in 80 µl of nuclease-free

water and stored at -80 ºC.

The RNA suspension was treated with RQ1 DNase (Promega, Madison, WI) to

remove contamination as follows. Twenty µl of 1:1 RQ1 DNase and 10X DNase buffer

was added to the RNA suspension. The mixture was incubated at 37 °C for one hr.

Following incubation, 200 µl nuclease-free water and 300 µl phenol (pH 4.3) were

added to the RNA suspension, vortexed, and centrifuged at high speed for 10 min. The

upper phase obtained after centrifugation was collected into a fresh tube. An equal

volume (0.6 mL) of phenol: chloroform was added to the contents in the tube, vortexed,

and centrifuged at high speed for 10 min. The upper phase was collected into a new

clean tube and sodium acetate precipitation was used to collect RNA (present in this

phase) as follows. One-tenth the volume of sodium acetate (3 M, pH 5.2), and 2.5

volumes of 100% ethanol were added to the upper phase collected and briefly inverted

to mix the contents. The RNA was precipitated by chilling overnight at -80 °C. The RNA

precipitate was pelleted by high speed centrifugation for 20 min and the supernatant

discarded. Three hundred µl of cold 70% ethanol were used to wash the pellet by

inverting the tube gently. The pellet was re-suspended in 50 µl of nuclease-free water.

A second clean-up of the RNA suspension was performed by chilling it for 30 min

at -20 °C in LiCl solution at a final concentration of 2.5 M. Stock LiCl solution was

prepared as 7.5 M in DEPC (1% diethylpyrocarbonate)-treated, nuclease-free water

(autoclaved) and filter sterilized with a 0.2 µm filter. The RNA (50 µl) was re-suspended

in nuclease-free water after pelleting by centrifugation at high speed (4 °C) and washing

with 70% ethanol. RNA concentration, purity, and integrity were verified via

74

spectroscopy using the NanoDrop™ 1000 (Thermo Scientific) instrument and running

on 1.2% agarose gel using DEPC-treated TAE (TAE- 40 mM tris, 20 mM acetic acid,

and 1 mM EDTA (pH 8) prepared with DEPC, treated water).

Quantitative Polymerase Chain Reaction (qPCR)

The qPCR reactions were performed using 150 ng of template DNA in 12.75 µl

total reaction volume with the SensiMix HRMTM Kits (Bioline, Taunton, MA). The

primers, BtCyt-f and BtCyt-r (Table 3-1) were used to amplify the Cyt2Ca1 gene, and

CitCD-f and CitCD-r were used to amplify the control gene (CD- citrus dehydrin- used

as a normalization control). The Cyt2Ca1 amplification was done in triplicate and CD in

duplicate for each transgenic citrus. The CAS1200 Liquid Handling System (Corbett

Robotics- now Qiagen, Valencia, CA) was used to aliquot multiple reactions

simultaneously, and the Rotor-Gene 6000 (Corbett Robotics) was used to run qPCR

reactions. The following thermal cycling conditions were used: 95 °C for 10 min,

followed by 40 cycles of 95 °C for 15 sec, 60 °C for 30 sec and 72 °C for 30 sec.

Amplification reactions were immediately followed by a melt curve analysis of the

product between 70-90 °C. The presence of amplified product was assessed based on

the number of amplification cycles needed to reach a common fixed threshold (cycle

threshold - Ct) in the exponential phase of PCR.

Quantitative Reverse Polymerase Chain Reaction (q-RT-PCR)

The QuantiTect® SYBR® Green (Qiagen) kit was used to conduct q-RT-PCR

reactions. The reactions were performed according to the manufacturer’s

recommendations but with half the final total volume (12.5 µl) suggested. To normalize

Cyt2Ca1 expression data, an experiment was carried out to identify a reference gene

that was stably expressed with the least variation across the root tissues. Three

75

reference genes, ubiquitin, actin, and 18S RNA, were assayed using RNA extracted

from 12 Cyt2Ca1 transgenic citrus and 1 non-engineered citrus control plant. An amount

of 100 ng of RNA templates was used per reaction with the primer pairs BtCyt-f and

BtCyt-r, CitUbi-f and CitUbi-r, and CitAct-f and CitAct-r to amplify Cyt2Ca1, ubiquitin,

and actin, respectively. However, 25 pg of RNA template were used in the 18S RNA

amplification reactions (primers: Cit18s-f and Cit18s-r). Primers were used at a final

concentration of 1 µM. Reverse transcription reactions were performed in duplicate

alongside single no RT (reverse transcriptase) controls to check for DNA contamination,

and no template controls with nuclease-free water to check for cross-contamination.

The CAS1200 Liquid Handling System was used to perform multiple reactions

simultaneously and the Rotor-Gene 6000 was used to run the reactions. The following

thermal cycling conditions were used: 50 °C for 30 min, 95 °C for 15 min, followed by 35

cycles of 94 °C for 15 sec, 60 °C for 30 sec and 72 °C for 30 sec. Amplification

reactions were immediately followed by a melt curve analysis from 70-90 °C to verify the

product. Expression levels were assessed based on the number of amplification cycles

needed to reach a common fixed threshold (cycle threshold - Ct) in the exponential

phase of PCR. Relative levels of transcription of Cyt2Ca1 gene in transgenic citrus

plants was normalized with the 18S RNA levels as the difference between the Ct of 18S

rRNA and Ct of Cyt2Ca1.

Potted-Plant Bioassays

An ad hoc criterion was used to classify transgenic citrus as high, medium, or low

expressers: Ct Difference greater than one = high expressers; Ct Difference less than

one, but greater than minus one = medium expressers; and Ct difference less than

minus one = low expressers (Figure 3-2). Cuttings were made from these plants (batch-

76

3). Growth of cuttings was observed to be uneven among the cuttings. Thus, plants that

grew to fairly uniform height were used for the potted plant bioassays. Seven Cyt2Ca1

transgenic citrus plants were selected for this experiment: low expressers (L-A3, L-422);

medium expressers (M-413, M-414); and high expressers (H-3A, H-U2, H-417). Using

an experimental design based on Lapointe and Bowman (2002), 10 cohorts of each

selected transgenic citrus (from 8-month-old cuttings) in addition to 10 cohorts of the

wild type (WT) Carrizo control were used for the potted plant bioassays. Figure 3-1

shows the experimental design.

The aim of the experiment was to challenge these transgenic citrus plants with

three-week-old D. abbreviatus larvae to determine their resistance to wounding as a

result of feeding injury and also to assess the effect on weight gain and mortality of

larvae that feed on transgenic plant roots. Citrus plants (10 each of transgenic citrus

and wild type controls) initially grown in yellow cones were transplanted into 2-L pots

(approximately 11 cm diameter, 25 cm tall) with drain holes containing white, sterile

sand. To prevent escape of larvae from infested pots, two pots were nested into each

other with a plastic screen placed between them and covering drain holes. Root-

volumes of each plant were measured by the displacement of water method before

planting into sterile soil as follows. Plant roots were individually placed into a volumetric

cylinder containing a known volume of water at a predetermined point on the stem (the

soil level). The volume of water displaced was recorded as the equivalent of the volume

of plant roots. After transplanting into a pot containing sand, plants were placed on

benches in the greenhouse and maintained on a fertilizer mix (N: P: K, 20:10:20) diluted

in water provided at a weekly rate of 150 mg/L. Temperature during the experimental

77

period fluctuated between a maximum of 35 °C and minimum of 15 °C, and relative

humidity varied between 34% and 93%. No supplemental light was provided.

Three-week-old D. abbreviatus larvae were obtained from the colony maintained

at USDA-USHRL and reared as described by Lapointe and Shapiro (1999). Forty plants

were infested with larvae as well as five pots containing sterile soil only. Non-infested

plant controls as shown in Figure 3-1 were also set up. In the infested treatments, six

larvae were weighed (20-35 mg individual weight), then placed individually in 10-cm

deep holes in the soil at a distance of 2 cm from the stem of the plant. Larvae and roots

were recovered 28 days later. The final root volume for each plant was determined by

the displacement of water method. The change in root-volume for each plant was

calculated as the ratio of the final root volume to the initial root volume. From this, an

average was determined for each uninfested cohort as the uninfested root ratio (URR).

Similarly, the infested root ratio (IRR) was calculated for infested potted plants. A root

damage index (RDI) was calculated for each citrus plant as the difference between URR

and IRR. Mortality and weights of live larvae were measured. Abbott’s formula (1925)

was used to adjust for mortality not associated with treatment (Cyt2Ca1 expressed in

citrus) to obtain the corrected mortality by subtracting mortality of larvae feeding on the

control wild type citrus roots from the mortality of larvae feeding on transgenic roots.

Corrected percentage mortality was transformed with the arcsine transformation to

normalize the data for statistical analysis. Data were subjected to statistical analysis

using GraphPad Prism 5 (La Jolla, CA) with Dunnett’s multiple comparison tests to

compare differences between transgenic citrus plants and wild type citrus control.

78

Results

Cyt2Ca1-Positive Transgenic Plants A total of 75 Cyt2Ca1 citrus plants (batch 1)

were verified for the presence of the Cyt2Ca1 construct. Of these, 31 transgenic citrus

plants were validated positive for the insert from qPCR analysis. To be considered

positive for Cyt2Ca1 construct, two out of three Cyt2Ca1 qPCR reactions had to have

amplified DNA product as determined by their Ct values and results of melt analysis

(data not shown). Trees representing these Cyt2Ca1-positive transgenic citrus plants

were selected for q-RT-PCR studies.

Cyt2Ca1 expression levels were assessed in the 31 transgenic citrus plants from

batch-2 (Table 3-2). The stability in the expression of ubiquitin and actin, and the

proportion of 18S rRNA in total RNA were assessed for suitability to correct for sample

to sample variation and to normalize Cyt2Ca1 expression data. Cycle threshold (Ct) of

Cyt2Ca1, 18S rRNA, ubiquitin, and actin were transformed to concentration and then

their means determined. The Fishers test was then used to compare differences in

variance between Cyt2Ca1 to 18S rRNA, ubiquitin, and actin. The variance between

Cyt2Ca1 was significantly different from ubiquitin, and actin (F=67050, df=11, P< 0.0001

and F=2069, df=11, P<0.001, respectively), but not 18S rRNA (F=2.38, df=11, P>0.05).

Therefore, 18s rRNA was used to normalize Cyt2Ca1 expression levels in transgenic

citrus plants. Ubiquitin, actin, and 18S rRNA are recommended as reference genes in

citrus (Boava et al., 2011; Yan et al., 2012). Even though the 18S RNA is generally not

recommended as a reference gene when examining changes in the expression of a

gene due to a treatment, because of commonly observed imbalances in mRNA and

rRNA proportions (Vandesompele et al., 2002), this was not the case in this study.

Since, the objective in this study was to determine differences in Cyt2Ca1 expression

79

among plants, not changes in Cyt2Ca1 gene expression due to a treatment, a reference

gene in the strictest definition, was not necessary. Therefore, the 18S RNA proportion of

total RNA was sufficient in this case to normalize gene expression.

Potted-Plant Bioassays

The uninfested root ratio (URR) was significantly higher in the high expressers

than in the low and medium expressers (Figure 3-5a). The wild type (WT) URR had the

highest growth rate increasing threefold during the experimental period, but it was not

significantly different (P > 0.05) from the values for transgenic plants H-3A, H-U2 and H-

417 (URR: 2.34, 2.63 and 2.95, respectively) in the Dunnett’s multiple comparison tests

(Table 3-3). The intrinsic growth of L-A3, L-422, M-413, and M-414 were significantly

lower, with URRs of 2.21, 1.60, 1.29, and 1.44, respectively. The root ratio was higher in

the high expressers than the low and medium expressers, contrary to earlier

observations. The observation that the root growth was higher in the high expressers

was the opposite of what was observed when the cuttings were initially growing (Figure

3-3). This suggests that the effect of Cyt2Ca1 on plant growth is influenced by factors

that were not measured in this study.

The impact of root feeding by larvae after 28 days is manifested by the IRR

(Figure 3-5b). The infested root ratio (IRR) of all the transgenic citrus plants did not

statistically vary from the wild type control in the Dunnett’s multiple comparison test

(Table 3-4). This suggests that larval feeding impacted evenly on the roots of transgenic

and non-transgenic citrus. However, the IRR fails to take into account the intrinsic

growth of roots. The URR normalizes the impact of feeding injury.

To account for the intrinsic root growth, a root damage index (RDI) was

calculated based on a simple assumption. If the Cyt2Ca1 toxin totally precludes larval

80

feeding on the roots, then there should be no significant difference between the URR

and the IRR for a treatment plant. The root damage index (RDI) is calculated as the

difference between URR and IRR. Ideally, an RDI close to zero is reflective of a citrus

plant that tolerates damage from larval root feeding (Figure 3-5c). A small RDI is

representative of resistance of roots to larval feeding damage. A large RDI however,

indicates a plant that is severely impacted by root feeding activity of larvae. In Figure 3-

4, it is observed that the damage to the root of the transgenic plant M-414 is less severe

than the wild type citrus plant. The caveat of the assumption in calculating the RDI is

that larval feeding could have been at the upper regions of the roots resulting in greater

root loss.

Results of the Dunnett’s multiple comparison tests of the RDI are shown in Table

3-5. Transgenic citrus event M-414 recorded the lowest root damage index with 0.50 ±

0.20 (Figure 3-5), but this was not significantly different from transgenic citrus events M-

413 (RDI = 0.61 ± 0.16) and L-422 (RDI = 1.06 ± 0.17) whereas the RDI of WT was

significantly higher (RDI of 2.53 ± 0.47). Larval feeding therefore had a greater impact

on the roots of the wild type control than on transgenic citrus events L-422, M-413 and

M-414 which are more resistant of larval feeding damage.

Weight gains (Figure 3-5d) of surviving larvae from all plant treatments were not

significantly different (Table 3-6). The highest larval mortality among the transgenic

citrus plants was recorded with event M-414 (Figure 3-5e), with a corrected mortality of

43.2 ±11.9%, significantly different (P < 0.10) from the wild type mortality of 8.4 ± 5.1%

(Table 3-7).

81

Discussion

Transformations to develop transgenic plants are not always successful (Clough

and Bent, 1998). Hence, methods of identifying successfully transformed plants are

necessary. In this study, quantitative polymerase chain reaction (qPCR) identified

transgenic citrus plants based on the presence of the Cyt2Ca1 synthetic gene in

regenerated plants from transformation experiments. Furthermore, methods of

assessing the levels of expression of the Cyt2Ca1 insert were necessary to identify

transgenic citrus plants for potted-plant bioassays. Polyacrylamide gel electrophoresis

and western blotting were used to assess the translation of Cyt2Ca1 but the data were

inconclusive. However via q-RT-PCR studies, the level of transcription of Cyt2Ca1was

assessed in cuttings from qPCR validated plants. I was able to develop an ad hoc

criterion (Figure 3-2) for the classification of 31 transgenic citrus plants into low,

medium, and high expressers, from analysis of the cycle threshold (Ct) data from q-RT-

PCR. With this criterion and an ability to achieve uniform growth, suitable transgenic

citrus plants were selected to test their root resistance to D. abbreviatus larval feeding

damage.

The expression of Cyt2Ca1 in the transgenic citrus influences root growth, as

indicated by the various uninfested root ratios (Figure 3-5a). The wild type URR was

significantly higher than three out of four of the evaluated transgenic citrus in the

bioassay (Table 3-3). This is contrary to the earlier observation made with the young

cuttings. High expression transgenic citrus plants developed at a slower rate than their

medium and low expressing cohorts (Figure 3-3). Perhaps this is an effect of age; once

there is a significant canopy available to support root growth, the energy redirected to

support high Bt expression may be insignificant to affect growth. Abiotic factors or

82

nutrition may have contributed to the values of URR attained. For ease of recapturing

larvae from the infested pots, and also to deny larvae other sources of nutrition, sterile

sand was used to set up the bioassays with a mixture of nitrogen, phosphorus, and

potassium provided as fertilizer. Perhaps a better nutritional regime can be designed for

future assays.

The larval mortalities in this study were not as high as those observed earlier by

Weathersbee et al. (2006). When neonates of D. abbreviatus were fed 300 µg/ml of

lyophilized, sporulated culture of Bacillus thuringiensis isolate expressing Cyt2Ca,

80.4% mortality was observed (Weathersbee et al., 2006). Perhaps larvae were

exposed to a lower dose of recombinant Cyt2Ca1 expressed in transgenic citrus roots.

Unfortunately, I was not able to quantify the Cyt2Ca1 recombinant protein. There are

possibilities to enhance the insecticidal activity of Cyt2Ca1 citrus. Synergism between

some of the known Cry and Cyt toxins exists (Pérez et al., 2005; Xue et al., 2005;

Sóberon et al., 2013). Therefore, other Cry toxins such as CryET33 and CryET43

(Weathersbee et al., 2006) may enhance the insecticidal activity of Cyt2Ca1. Peptide

moieties may also enhance the toxicity of Cyt2Ca1 expressed in citrus. The addition of

a short peptide to Cyt2Aa enhanced its toxicity to the aphids Acyrthosiphon pisum

Harris and Myzus persicae Sulzer (Chougule et al., 2013). Likewise, TMOF may

enhance the toxicity of Cyt2Ca1 in transformed citrus. The transgenic citrus plant M-414

may be a candidate for the addition of other toxins that work via other modes of action

to elicit a mortality response.

The root damage index (RDI) is useful in a study such as this because of the

intrinsic differences in root growth ratios. The RDI (Figure 3-5b) normalizes the impact

83

of feeding damage (Figure 3-5c) on the transgenic citrus roots and the wild type citrus

controls. The analysis of RDI culminated in the selection of transgenic plants L-422, M-

413, and M-414 for further field studies. Importantly, these transgenic citrus scored

significantly lower root damage indices (Table 3-5), satisfying the criterion set to identify

transgenic citrus plants resistant to root feeding damage.

A high mortality in D. abbreviatus larvae and a low RDI are two important

characteristics of a D. abbreviatus resistant plant. The transgenic citrus event M-414

had the best of traits amongst the transgenic citrus plants tested in the potted plant

bioassays. Plant M-414 was established to be a medium expresser of Cyt2Ca1 though

it is not clear now why it performed better than both low and high expressers in the

study.

84

Figure 3-1. Experimental design of potted plant bioassay. A total of 70 transgenic citrus

cohorts from 7 transgenic citrus plants and 10 wild type cohorts were used for the

bioassay. Data collected to provide evidence of resistance included: larval mortality,

changes in weight of larvae (mg), differences in root volume (mL) remaining on the plant

after the feeding trial.

85

Figure 3-2. Relative levels of transcription of Cyt2Ca1 gene in transgenic citrus plants normalized with the 18S RNA levels. Transgenic citrus were categorized into high expressers (Ct Difference ≥ 1), medium expressers (Ct Difference less than 1, but greater than -1), and low expressers (Ct Difference ≤ -1). Plants with red, green, and yellow columns were used for potted plant bioassays.

86

Figure 3-3. Some transgenic citrus cuttings used for potted plant bioassays. An observation was made that the rate of growth was inversely proportional to the level of transcription of recombinant Cyt2Ca1 gene; H.E.; high expresser, M.E.; medium expresser, L.E.; low expresser. Photo by S.K. Ben-Mahmoud,

L.E. M.E. H.E.

87

Figure 3-4. Resistance of Cyt2Ca1 engineered citrus versus wild type control. Transgenic citrus plant M-414 (a), non-engineered control citrus plants (b).The root mass excised in (b)- arrowed, is considered as damage/loss. Photos by S.K. Ben-Mahmoud.

(a) (b)

88

Figure 3-5. Effect of larval root feeding on Cyt2Ca1 citrus plants and the concomitant effect on larval weight and mortality. L-A3, L-422, M-413, M-414, H-3A, H-U2 and H-417 are transgenic citrus plants. WT is a wild type citrus control and SOIL is the starved control. Uninfested plant cohorts (a) were set up alongside infested plant cohorts (b) in which 6, 3-week old larvae fed on roots for 28 days (5 plants each). Changes in root volume were recorded as the infested root ratio (IRR) (b) and uninfested root ration (URR) (a). A root damage index (URR-IRR) was assessed from the infested and uninfested potted plants (c). Weight changes of surviving larvae (d) and mortality of larvae fed on plant roots in potted-plant bioassays are also shown (e). Bars sharing the same letter in each were significantly different from the wild type.

89

(a) Uninfested Root Ratio

L-A3

L-422

M-4

13

M-4

14

H-3

AH-U

2

H-4

17 WT

0

1

2

3

4

a

aa a

Citrus Plant

RO

OT

VO

LU

ME

RA

TIO

(D

ay28/D

ay0)

(b) Infested Root Ratio

L-A3

L-422

M-4

13

M-4

14

H-3

AH-U

2

H-4

17 WT

0.0

0.5

1.0

1.5

Citrus Plant

RO

OT

VO

LU

ME

RA

TIO

(D

ay28/D

ay0)

(c) Root Damage Index

L-A3

L-422

M-4

13

M-4

14

H-3

AH-U

2

H-4

17 WT

0

1

2

3

4

b

b b

Citrus Plant

UR

R-I

RR

(d) Weight Gain of Surviving Diaprepes fed on Cyt2Ca1 Citrus

L-A3

L-422

M-4

13

M-4

14

H-3

AH-U

2

H-4

17 WT

SOIL

0

50

100

150

200

c

Citrus Plant

AV

. W

EIG

HT

OF

LIV

E L

AR

VA

E (

mg

)

(e) Mortality of Diaprepes larvae fed on Cyt2Ca1 Citrus

L-A3

L-422

M-4

13

M-4

14

H-3

AH-U

2

H-4

17 WT

SOIL

0

20

40

60

80

d

d

d

Citrus Plant

% C

orr

ecte

d M

ort

ali

ty

90

Table 3-2. Summary statistics of Ct values obtained from q-RT-PCR analysis of Cyt2Ca1 citrus and non-engineered citrus

Cycle Threshold (Ct)

Cyt2Ca1 18S

rRNA Ubiquitin Actin

Mean 11.75 12.49 23.11 17.43

Standard Error 0.41 0.15 0.77 0.53

Standard Deviation 2.26 0.86 2.76 1.89

Sample Variance 5.09 0.75 7.63 3.59

Coefficient of Variation-CV (%) 7.28 2.70 21.25 14.57

Range 8.73 3.85 10.53 7.10

Minimum 9.11 9.83 19.39 15.69

Maximum 17.83 13.67 29.92 22.79

Count 31.00 32.00 13.00 13.00

91

Table 3-3. Statistical analysis of uninfested root ratio

Dunnett's Multiple Comparison Test Mean Diff. Q P < 0.05? Summary 95% CI of diff

WT versus L-A3 0.8980 3.006 Yes * 0.07233 to 1.724

WT versus L-422 1.508 5.048 Yes *** 0.6823 to 2.334

WT versus M-413 1.822 6.099 Yes *** 0.9963 to 2.648

WT versus M-414 1.668 5.584 Yes *** 0.8423 to 2.494

WT versus H-3A 0.7660 2.564 No Ns -0.05967 to 1.592

WT versus H-U2 0.4780 1.600 No Ns -0.3477 to 1.304

WT versus H-417 0.1600 0.5356 No Ns -0.6657 to 0.9857

Transgenic citrus plants: L-A3, L-422, M-413, M-414, H-3A, H-U2 and H-417. WT: Wild Type control.

Table 3-4. Statistical analysis of infested root ratio

Dunnett's Multiple Comparison Test Mean Diff. Q P < 0.05? Summary 95% CI of diff

WT versus L-A3 -0.04000 0.1678 No Ns -0.6987 to 0.6187

WT versus L-422 0.04000 0.1678 No Ns -0.6187 to 0.6987

WT versus M-413 -0.09800 0.4112 No Ns -0.7567 to 0.5607

WT versus M-414 -0.3600 1.511 No Ns -1.019 to 0.2987

WT versus H-3A -0.3020 1.267 No Ns -0.9607 to 0.3567

WT versus H-U2 -0.3480 1.460 No Ns -1.007 to 0.3107

WT versus H-417 0.0880 0.3693 No Ns -0.5707 to 0.7467

Transgenic citrus plants: L-A3, L-422, M-413, M-414, H-3A, H-U2 and H-417. WT: Wild Type control.

92

Table 3-5. Statistical analysis of root damage index

Dunnett's Multiple Comparison Test Mean Diff. Q P < 0.05? Summary 95% CI of diff

WT versus L-A3 0.9380 2.317 No Ns -0.1810 to 2.057

WT versus L-422 1.468 3.626 Yes ** 0.3490 to 2.587

WT versus M-413 1.920 4.742 Yes *** 0.8010 to 3.039

WT versus M-414 2.024 4.999 Yes *** 0.9050 to 3.143

WT versus H-3A 1.070 2.643 No Ns -0.04900 to 2.189

WT versus H-U2 0.8280 2.045 No Ns -0.2910 to 1.947

WT versus H-417 0.07000 0.1729 No Ns -1.049 to 1.189

Transgenic citrus plants: L-A3, L-422, M-413, M-414, H-3A, H-U2 and H-417. WT: Wild Type control.

Table 3-6. Statistical analysis of weight gain of surviving larvae fed on citrus

Dunnett's Multiple Comparison Test Mean Diff. Q P < 0.05? Summary 95% CI of diff

WT versus L-A3 -4.630 0.1450 No ns -93.91 to 84.65

WT versus L-422 30.74 1.021 No ns -53.44 to 114.9

WT versus M-413 56.48 1.876 No ns -27.70 to 140.7

WT versus M-414 51.54 1.712 No ns -32.64 to 135.7

WT versus H-3A 61.42 2.040 No ns -22.76 to 145.6

WT versus H-U2 -18.70 0.6210 No ns -102.9 to 65.48

WT versus H-417 -13.26 0.4404 No ns -97.44 to 70.92

WT versus SOIL 111.4 3.700 Yes ** 27.24 to 195.6

Transgenic citrus plants: L-A3, L-422, M-413, M-414, H-3A, H-U2 and H-417. WT: Wild Type control.

93

Table 3-7. Statistical analysis of larval mortality

Dunnett's Multiple Comparison Test Mean Diff. Q P < 0.10? Summary 95% CI of diff

WT versus L-A3 -0.1974 1.365 No Ns -0.6011 to 0.2063

WT versus L-422 -0.2081 1.439 No Ns -0.6118 to 0.1956

WT versus M-413 -0.2219 1.535 Yes * -0.6257 to 0.1818

WT versus M-414 -0.3769 2.606 Yes ** -0.7806 to 0.02681

WT versus H-3A -0.1974 1.365 Yes * -0.6011 to 0.2063

WT versus H-U2 0.04241 0.2932 No Ns -0.3613 to 0.4461

WT versus H-417 0.0000 0.0000 No Ns -0.4037 to 0.4037

WT versus SOIL -0.6096 4.214 Yes ** -1.013 to -0.2058

Transgenic citrus plants: L-A3, L-422, M-413, M-414, H-3A, H-U2 and H-417. WT: Wild Type control.

94

CHAPTER 4 TARGETING THE PERITROPHIC MEMBRANE OF DIAPREPES ABBREVIATUS VIA

RNAi INDUCES A DOSE-DEPENDENT MORTALITY RESPONSE

Introduction

RNA interference (RNAi) is a recently discovered mechanism in eukaryotes

(Shabalina and Koonin, 2008). RNAi is an important defense against viruses and

transposable elements (Obbard et al., 2009), but the ability to manipulate it for pest

control has been shown (Baum et al., 2007; Mao et al., 2007). RNAi is a sequence-

specific, post-transcriptional silencing of genes, as a result of the degradation of target

mRNA, such that the ability of a cell to produce the cognate protein is greatly diminished

(the cognate gene is said to be silenced). The common underlying theme of RNAi is the

requirement of a double-stranded RNA (dsRNA) template molecule (Meister and

Tuschl, 2004). In RNAi for pest control, the template is typically a relatively large dsRNA

(typically between 200 and 600 bps).

Armed with the sequence data of organisms, it is possible to design and

synthesize dsRNA molecules for the purpose of silencing genes important for

physiological functions within a target organism. These genes are essential genes

without which organisms have reduced fitness or cannot survive. In insects, these

genes might be involved in homeostasis: the insulin signaling and forkhead transcription

factor (FOXO) silenced in Culex pipiens Linnaeus (Sim and Denlinger, 2008), in

digestion: the digestive cellulose enzyme in Reticulitermis flavipes (Kollar) (Zhou et al.,

2007), and in metabolism: cytochrome P450 in Helicoverpa armigera (Hübner) (Mao et

al., 2007) and Plutella xylostella (Linnaeus) (Bautista et al., 2009). Other examples

include genes involved in growth and development: Tc-achaete-scute-homolog (Tc-

ASH) in Tribolium castaneum (Herbst) (Tomoyasu and Denell, 2004), in reproduction

95

and caste differentiation in termites: hexamerin genes in termites (Zhou et al., 2006),

and in blueprints for structural proteins, such as α-tubulin in Diabrotica virgifera

(LeConte) (Baum et al., 2007).

RNAi as a pest control tool is novel, but the pace of its development is

precocious. Although not commercially available yet, the feasibility of this tool as a pest

management strategy has already been shown demonstrated and is now hotly debated

(Gatehouse, 2008). The specificity in the design of dsRNA that can be achieved is often

cited as its biggest advantage. However, issues with quality assurance exist as non-

target effects have also been observed that are yet to be fully understood.

In insects, two inherent characteristics are to be considered in the selection of

target genes. First, barriers such as the peritrophic membrane and cuticle, in addition to

potential degradation of dsRNA molecules, may impair the uptake of dsRNA molecules

(Terenius et al., 2011; Bachman et al., 2013). Second, amplification of the dsRNA

template/signal may be absent. As yet, the enzyme RNA-dependent RNA polymerase

(RdRP) which is responsible for RNAi signal amplification, or its homologues, have not

been found in any insect (Tomoyasu et al., 2008; Duan et al., 2013), though systemic

spread of dsRNA has been shown in some insects (Terenius et al., 2011; Knorr et al.,

2013). These two characteristics make the selection of targets important because the

dsRNA molecules must be able to reach their target tissues to elicit a response.

Delivery of dsRNA molecules to insects in laboratory studies has mainly involved

microinjections, oral uptake as part of artificial diets and transgenic plant varieties, and

topical applications in a mixture with cuticular penetrants. In commercial application as a

strategy for pest insect control, the delivery method of choice, for any pest, will depend

96

on factors such as cost of dsRNA synthesis, feeding behavior and life cycle of pest.

Most likely, there will not be a “one-size-fits-all” approach to the adoption of this

technology. For field pest control purposes, the oral uptake route is likely to be the most

pragmatic. It is envisioned that this would be commercialized by expressing double-

stranded RNA in transgenic plants or by expression in crop plants using an attenuated

plant virus delivery strategy.

Larvae of D. abbreviatus are the most damaging stage of the pest as a result of

root feeding injuries to citrus trees. Novel strategies are needed because no effective

pest management options are available. RNAi might be a suitable tactic to complement

peptide inhibitors of digestion and Cyt2Ca1 citrus for D. abbreviatus control in a novel

transgenic strategy. It will not be pragmatic to deliver dsRNA molecules as spray

formulations because D. abbreviatus larvae, the most destructive stage of the weevil

pest, is effectively, entirely subterranean. Additionally, it is not yet economical to

produce dsRNAs in quantities that will allow soil drenching, assuming that the dsRNAs

are totally innocuous to the environment and non-target organisms. Transgenic citrus or

attenuated CTV virus, expressing dsRNA that work to silence D. abbreviatus specific

genes will be a feasible approach in a manner similar to Baum et al. (2007). Transgenic

corn engineered to express D. virgifera α-tubulin V-ATPase A dsRNA were protected

from larval feeding damage. Since larvae feed on the roots of citrus, genes expressed in

the gut might be more appropriate targets due to direct access to potentially vulnerable

tissues. To this end, the objective of this study is to identify targets that will achieve

significant mortality for an RNAi strategy for D. abbreviatus control.

97

Materials and Methods

Diaprepes abbreviatus Larvae

Two- to three-week-old D. abbreviatus larvae maintained on an artificial diet

(Lapointe and Shapiro, 1999) were obtained from a colony at the USDA-USHRL. Larvae

were individually weighed to select those between 15-30 mg for use in larval bioassays.

Larvae were randomly assigned to treatments.

Primers

Primers were designed and purchased from IDT® (San Diego, CA) to amplify

approximately 300 base pair regions within D. abbreviatus for alpha-tubulin, peritrophic

matrix protein, peritrophic membrane chitin binding protein, and cathepsin L2. The

primer sequences are listed in Table 4-1 in pairs (forward and reverse). The T7

promoter sequences (highlighted in Table 4-1) were added to primers that were used for

synthesizing dsRNA molecules. The melting temperatures are also listed as they guided

the cycling conditions of polymerase chain reactions.

Double-stranded RNA Synthesis.

At the outset, cDNA clones of the target genes were made and stored in the

vector, pCR®2.1-TOPO® (Invitrogen, Grand Island, NY). Approximately 300 base pair

regions of the target genes (α-tubulin, peritrophic matrix protein, peritrophic membrane

chitin binding protein, and cathepsin L2) were identified for dsRNA synthesis. For

control treatments, a random dsRNA sequence from the Asian citrus psyllid (Diaphorina

citri) was obtained from Dr. R. Shatters’ laboratory at the USDA-USHRL. Template DNA

for dsRNA synthesis was synthesized with standard PCR reactions using the AmpliTaq

Gold® kit (Roche, Indianapolis, IN) according to the manufacturer’s protocol. Since all

98

the primers had high melting temperatures (greater than 70 °C), the touchdown PCR

cycling conditions were selected.

Table 4-1. Primers for Diaprepes abbreviatus dsRNA synthesis. Target Primer Name

(forward/Reverse) Sequence (5’- 3’) Tm (melting

temperature) °C

α-Tub DB 1108 TAATCAGACTCACTATAGGGGACTCTGAGTAATATAGTCAACTAAAGCAG

73.7

DB 1109 TAATCAGACTCACTATAGGGCCAAGTCTACGAATACCGCTCTGG

79.2

PMP DB 1126 TAATCAGACTCACTATAGGGGAGTCTTCAGTTGAAAAATCTTTGAAA

75.8

DB 1127 TAATCAGACTCACTATAGGGGTAGTGACAGGTGGCCTAACGGTA

78.4

PMCBP DB 1102 TAATCAGACTCACTATAGGGCATGAAGGTTCACGTTATTCTTTG

76.5

DB 1103 TAATCAGACTCACTATAGGGTGATAGTCACTGTATTCATGTGGTACA

75.1

CAL 2 DB 1118 TAATCAGACTCACTATAGGGATGTATAGTTTAGTA

GTGTTGTGGCA 73.6

DB 1119 TAATCAGACTCACTATAGGGAGGTAATTGAGGCATATTTACTGTGTA

74.0

Legend: α-Tub; alpha-tubulin, PMP; peritrophic matrix protein, PMCBP; peritrophic membrane chitin binding protein, CAL2; cathepsin L2. T7 promoter sequences are highlighted.

The PCR product was quantified (NanoDrop™ 1000 –ThermoScientific, Waltham, MA)

and run on 1.2% agarose gels to verify purity of the amplification product. Amplified

products were purified with the QIAquick PCR purification kit (Qiagen, Valencia, CA)

using the recommended protocol. The HiScribeTM in vitro transcription kit (New England

Biolabs®, Ipswich, MA) was used for dsRNA synthesis. Nuclease-free water was used

for all RNA synthesis. Amount of template (PCR product) used per 40-µl reaction was

approximately 1 µg, and reactions were incubated at 42 °C overnight. One percent

agarose gels (TAE- 40 mM Tris, 20 mM acetic acid, and 1 mM EDTA (pH 8) prepared

with DEPC -treated water were used to check for the integrity of the dsRNA product.

Ammonium acetate precipitation was used to clean the transcription product as follows:

One-tenth volume of 5 M ammonium acetate and 2 volumes of cold ethanol (100%)

were added to the dsRNA product, mixed thoroughly by vortexing for 5 sec, and chilled

99

at -80 °C. Centrifugation at high speed (10 min, 4 °C) was used to create a pellet of

dsRNA. The supernatant was decanted and the pellet washed with 70% ethanol, after

which a pellet was collected via centrifugation (as before) and then briefly air-dried.

Double-stranded RNA was reconstituted in 200 µl RNAse-free water and the

concentration and purity were verified via spectroscopy (NanoDrop™ 1000 -Thermo

Scientific). All dsRNA products were stored at -80 °C until ready for use in larval

bioassays.

Larval Bioassay

A bioassay method, developed to test individual neonate larvae (Weathersbee et

al., 2006), was adopted with clear, 0.2-mL polymerase chain reaction (PCR) tubes to

verify tolerance to dsRNA molecules. D. abbreviatus specific dsRNA molecules and a

control D. citri dsRNA sequence were fed to larvae in artificial diet bioassays as

described herein. The D. citri dsRNA sequence served as a positive control for off-target

effect of non-specific sequences on D. abbreviatus RNAi machinery. Stock of each

dsRNA was diluted with nuclease-free water to the following concentrations: 0.05, 0.20,

0.50, and 1.00 µg/µl water. Fifty-µl aliquots of artificial diet (Lapointe and Shapiro, 1999)

were dispensed into the lids of individual PCR tubes and allowed to cool to room

temperature. Ten µl of one of each dsRNA solution concentration or D. citri dsRNA were

evenly applied to the surface of the artificial diet in a lid and allowed to percolate to yield

the following dosage treatments: 0.5, 2.0, 5.0, and 10 µg per 50 µl artificial diet for each

dsRNA. Each treatment was replicated 20 times. Tubes were air-dried for approximately

20 min in a sterile environment, and individual larvae were randomly transferred per

tube. Three control treatments consisted of 1) individually starved larvae and 2) artificial

diet without dsRNA with 10 µl nuclease free water, and 3) artificial diet only (to correct

100

for no treatment mortality). For each control treatment, 30 larvae were assayed. The

tubes were turned upside down in a sterile centrifuge tube box and kept for 12 days in

the dark in an environmentally controlled chamber set at 26 °C. Larval mortality was

assessed on the twelfth day.

Percentage mortality was corrected with the mortality from diet only, using the

Abbott’s (1925) formula. Statistical analysis of mortality data was done with the

GraphPad Prism 5 (La Jolla, CA). Percentage mortality was transformed with the

arcsine transformation to normalize data for statistical analysis. Differences among

treatment level means were determined by Bonferroni’s multiple comparisons tests to

compare D. abbreviatus dsRNA sequence mortality to the diet and water control

mortality.

Results

There was 30% mortality in larvae that fed on diet only 12 days post feeding.

After percentage mortality (± SE) was corrected for no treatment deaths, 4.8 ± 3.0% of

larvae died due to starvation (Figure 4-1), confirming the ability of D. abbreviatus larvae

to survive for long period without food. Mortality due to water added to diet resulted in

two times the mortality in starved larvae, but was not significantly different (t value= 2.2,

P>0.05). It is not unusual that the D. citri dsRNA resulted in non-target mortality in D.

abbreviatus mortality (Kulkarni et al., 2006; Nunes et al., 2013), but it is curious that the

mortality was not dose-dependent. Mortality as a result of D. citri dsRNA was not

significantly different from the diet and water control (t value 3.27, P>0.05). Larval

mortality as a result of dsRNA sequences targeting specific D. abbreviatus genes

caused significantly (P<0.05) higher mortalities than the control, ranging from 14.3 ±

4.5% to 71.4 ± 7%.

101

At the highest dose of dsRNA fed to larvae, the peritrophic membrane chitin

binding protein- PMCBP resulted in the significantly highest mortality (t value= 3.2,

P<0.05), and at 5.0 µl dsRNA/ 50 µl diet the cathepsin L2- CAL2 dsRNA caused the

significantly highest mortalities (t value= 4.9, P<0.01). However, at 2.0 µg dsRNA/ 50 µl

diet none of the D. abbreviatus dsRNA sequences were significantly different from the

D. citri control mortality (P>0.05), although the CAL2 and α-Tub dsRNA sequences

caused the highest mortalities. At 0.5 µg dsRNA/ 50 µl diet, alpha-tubulin caused the

significantly highest mortality (t value= 4.9, P<0.001).

Importantly, two of the D. abbreviatus dsRNA sequences, PMCBP and CAL2

resulted in mortality that was dose-dependent; the higher the dose the greater the

mortality. The PMCBP dsRNA resulted in 71.4 ± 7.0%, 42.9 ± 9.5%, 35.0 ± 4.0% and

14.3 ± 4.5% at 10.0, 5.0, 2.0 and 0.5 µg/ 50 µl diet respectively, whereas CAL2 resulted

in 50.0 ± 7.0%, 57.1 ± 9.5%, 42.9 ± 4.0% and 28.6 ± 4.5% with decreasing dose of

dsRNA. Mortality due to PMCBP was significantly different at 10 and 5 µg dsRNA / 50

µL artificial diet (P<0.01) whereas cathepsin L2 was significantly different at 5 (P<0.01)

and 0.5 µg dsRNA / 50 µL artificial diet (P<0.05). Mortality due to alpha-tubulin and

peritrophic matrix protein dsRNA sequences was not dose-dependent. The highest

mortality for α-tubulin was 57.1 ± 7.0%, whereas that for the peritrophic matrix protein

was 42.9 ± 7.0%.

Discussion

Feeding D. citri dsRNA to D. abbreviatus larvae elicited mortality significantly

higher than what was observed in diet without dsRNA. It is interesting that this non-

specific response was not dose dependent. Perhaps, dsRNA molecules beyond a

certain threshold induce some cellular response, unknown at this point that generally

102

affects global gene expression leading to death. In a study by Nunes et al. (2013), a

green fluorescent protein (GFP)-derived dsRNA sequence affected the expression of

about 10% of the genes of the honey bee, Apis mellifera, even though this gene is not

present in the bee. In the processing of dsRNA fragments to siRNA fragments, several

sequences may be generated that share similarities to non-specific genes and therefore

may affect their expression. This is likely the case with the D. citiri dsRNA sequence

used here, however it is not clear why this effect was not dose-dependent.

Two dsRNA sequences, PMBCP and CAL2, targeting the peritrophic membrane

and the digestive enzyme cathepsin L2, respectively, are good candidates for a novel

strategy for D. abbreviatus control by exploiting RNAi. They elicit mortality that is dose-

dependent in larvae (Figure 4-1). The peritrophic membrane plays key roles in insects,

including the compartmentalization of digestion to improve efficiency (Terra, 2001) and

serving as a physical barrier to prevent mechanical damage (Rao et al., 2004) and to

protect against invading and infectious microbial agents such as protozoa (Coutinho-

Abreu et al., 2013). Cathepsins, meanwhile, are one of two, important digestive

enzymes needed for growth and development of D. abbreviatus (Yan et al., 1999).

Significantly higher mortalities elicited as a result of feeding dsRNA sequences targeting

genes such as PMCBP and CAL2 involved in larval physiology and biochemistry of D.

abbreviatus is indicative of the relevance of the peritrophic membrane and cathepsin L2

to D. abbreviatus’ survival.

The RNAi response in insects is typically specific (Baum et al., 2007; Mao et al.,

2007), but non-target and off-target effects also occur (Zhang et al., 2010; Nunes et al.,

2013). These unintended effects are the result of short-interfering RNA (siRNA)

103

sequence similarities to non-target mRNAs when it is processed from dsRNA by dicer

(Qiu et al., 2005). Also, the saturation of the RNAi machinery by non-specific sequences

(Nunes et al., 2013) may overwhelm the organism. It is therefore very significant that

the dsRNA sequence targeting the peritrophic membrane elicited larval mortality that

correlates with the dosage administered (Figure 4-1). This consistency is relevant

towards establishing the validity of an RNAi effect (Terenius et al., 2011).

In spite of the mortality effects induced by dsRNA targeting alpha tubulin and

peritrophic matrix protein, they are not useful candidates for RNAi in D. abbreviatus due

to the inconsistencies with dose (Figure 4-1). However, this is not to imply that the

targets selected are poor. Perhaps other sequence regions may be better at silencing

the respective targets. The rational design of effective sequences is currently very

challenging. A number of sequences, though not entirely random, are typically screened

and fortuitously a candidate sequence is obtained. Perhaps new computational

predictors (Sciabola et al., 2013) may improve currently identified effective sequences

to be more viable.

As a result of the limitations of exploiting RNAi in insects, genes relevant for

survival, growth, and development and expressed in the midgut are the choice targets.

So far, the peritrophic membrane chitin binding protein particularly and cathepsin L2

have shown to be amenable targets in a future pest control strategy in D. abbreviatus.

104

-10.0

0.0

10.0

20.0

30.0

40.0

50.0

60.0

70.0

80.0

90.0

10 µg/diet 5 µg/diet 2 µg/diet 0.5 µg/diet

Psyllid dsRNA

α-Tub

PMP

PMCBP

DaCatL

Diet + water

Figure 4-1. Corrected mortality (post 12 days) due to feeding on dsRNA sequences targeting Diaprepes abbreviatus genes. Legend: control dsRNA sequence targeting a Diaphorina citri gene (Psyllid dsRNA); alpha-tubulin (α-Tub), peritrophic matrix protein (PMP), peritrophic membrane chitin binding protein (PMCBP), cathepsin L2 (CAL2), of D. abbreviatus.

105

CHAPTER 5 GENERAL CONCLUSIONS AND FUTURE DIRECTIONS

Toxins with different modes of action are particularly desirable for new transgenic

crop varieties. The single-toxin Bt crop varieties have been commercially deployed in

the field for over a decade, and have exhibited remarkable resilience against resistance

selection in insect populations (Bates et al., 2005). However, it has become apparent

that they may not be sustainable any longer. Indeed, there are recent reports of field-

evolved resistance in insects to the single-toxin Bt crops (Liu et al., 2010; Gassmann et

al., 2011). Multiple-toxin transgenic plants are expected to further delay the inception of

resistance. Thus, the ultimate aim for a transgenic strategy in citriculture is to have

citrus express multiple toxins that act through different modes leading to death of larvae

that feed on roots. Candidate toxins were explored in a three-pronged approach aimed

at inhibiting digestive enzymes (cathepsins) with peptide inhibiters in in vitro enzyme

assays, disrupting gut integrity with the Bt toxin Cyt2Ca1 expressed in transgenic citrus,

and interference with gene expression via RNAi by feeding dsRNA molecules in artificial

diet bioassays.

Protease inhibitors originating from plants have been investigated for control

purposes. Some level of protection has been demonstrated in certain instances, but the

majority of results have been mediocre; indeed, in some cases the opposite effect has

been observed (Murdock and Shade, 2002). Insects are highly adaptable and have

circumvented inhibitors simply through an apparent increase in the expression of

digestive enzymes (Adbeen et al., 2005) or switching to a different complement of

proteases or sequestration of the inhibitor or all of the above. In a stacking/pyramiding

strategy, inhibiting the activity of digestive cathepsins with CPIs may be enough to make

106

the insect more vulnerable to additional toxins, such as Cyt2Ca1, despite their best

efforts at detoxifying the inhibitor. The availability of pure recombinant enzyme enables

the specific determination of the binding characteristics of the enzyme and inhibitor. I

set forth to characterize the interaction between DaCatL1 and CPI1.

My attempts at synthesizing and purifying a Diaprepes abbreviatus cathepsin for

inhibitor characterization with CPI resulted in the serendipitous discovery of a basic

cathepsin, DaCatL1, expressed in E. coli. Recombinant DaCatL1 was only stable in

solution with 1% TritonTM X-100; DaCatL1 was therefore folded favorably in TritonTM X-

100. Whether TritonTM X-100 has any effect on the chemical characteristics of DaCatL1

remains to be seen. Sequence comparison of DaCatL1 to other cathepsins revealed the

same amino acid residues at the active sites. Protein folding is apparently important to

their activity and chemical characteristics. The chemical characteristics of DaCatL1 are

cryptic. DaCatL1 has optimum activity at pH 8 with the general cathepsin L substrate Z-

FR-AFC. Whereas DaCatL1 activity was inhibited by chymostatin, a general protease

inhibitor with activity against cathepsins, it was not inhibited by the specific cathepsin L

inhibitors E-64 and Z-FY(tBu)DMK. Failure to inhibit DaCatL1 with the specific cathepsin

L inhibitors may be explained by the pH optima of the enzyme. Importantly, aprotinin, a

serine protease inhibitor, does not inhibit DaCatL1 at the same concentration effective

with the serine proteases which have pH optima in the basic range. X-ray

crystallographic studies may be utilized to resolve this conundrum.

A recombinant CPI1, an endogenous modulator of cathepsin in D. abbreviatus,

and a mutant CPI1 (CPI1mut) were purified to near homogeneity. CPI1 mutant has

Lys101 replaced with a cysteine residue, making it a more robust inhibitor of D.

107

abbreviatus cathepsins. The docking models of CPI1 and DaCatL1 reveal proximity of

Lys101 to the active site residue Cys120 of DaCatL1. Thus, mutating Lys101 to a Cys

was hypothesized to result in the formation of a stronger S-S bond which will be

unaffected by pH. In this way, the interaction of CPI1 with cathepsin will be changed

from a reversible interaction to an irreversible one. I was unable to show CPI1mut to be

a better inhibitor of D. abbreviatus gut cathepsins across the acidic to basic range.

However, CPI1 and CPI1mut are both effective inhibitors of acidic cathepsins. Added to

the inability to inhibit the basic gut cathepsins, CPI1 and CPI1mut were weak inhibitors

of DaCatL1. This sheds new insights into the regulation of digestive cathepsins in D.

abbreviatus. Two activity peaks of gut cathepsins were obtained, one acidic and the

other basic. Therefore, the acidic and basic cathepsins appear to be uniquely distinctive

and may have different modulators of their activity.

Transgenic citrus engineered to express Cyt2Ca1 needed to be evaluated.

Although, the exact mode of action of Cyt2Ca1 was not determined in this study, it was

expected to insert into gut epithelial membranes and disrupt gut integrity resulting in

death. Importantly, three transgenic citrus plants resisted feeding injuries by D.

abbreviatus larva. Using a root damage index, these three plants were less impacted by

feeding damage. Additionally, one of the transgenic plants caused significantly high

mortality to D. abbreviatus larvae. Choice tests and/or field studies might further reveal

the resilience of Cyt2Ca1 citrus to D. abbreviatus larval feeding damage. However,

larval mortality, and weight losses, of three-week old D. abbreviatus when fed on

transgenic citrus were low or not substantial. There may be possibilities of enhancing

Cyt2Ca1 insecticidal activity by adding small peptide moieties.

108

The interruption of gene expression via RNAi has become a new tool in the

development of pest control tactics. Preliminary studies investigating the expression of

D. abbreviatus genes exposed to dsRNA sequences in artificial diet show the possibility

of exploiting RNAi for citrus protection. Sequences of 300 base pairs designed to

hybridize with the peritrophic membrane chitin binding protein (PMCBP) and cathepsin

L2 (CAL2) mRNAs elicited dose-response mortality when fed to 2-3- week-old D.

abbreviatus larvae. The specificity of this response needs to be validated because non-

target effects cannot be ruled out entirely at this point. The dsRNA sequences could be

engineered directly in transgenic citrus or could be expressed by an attenuated citrus

virus vector. Currently, a CTV vector is being engineered by scientists at the University

of Florida Citrus Research and Education Center (Lake Alfred, FL) to express the

PMCBP sequence.

Whereas transgenic citrus with its own armor against D. abbreviatus represents

a potential pest management tactic, there are potential bottlenecks to its adoption.

Despite the rapid adoption of transgenic crops worldwide and greater public acceptance

of their products, there are still some concerns. From experience in applying the

transgenic technology in other commercial crops, impediments to faster development

and adoption include cost, regulations (Bradford et al., 2005), public perceptions of the

technology (Rommens, 2010), and concerns related to the safety of the foods and the

environment (Shelton et al., 2002). Other issues often raised include the risk of gene-

flow from transgenic plants to non-transgenic plants (DiFazio et al., 2012), biological

fitness of the transgenic plant (Gassmann et al., 2009), allergenicity of the transgenic

109

product (Goodman et al., 2008), and resistance development (Tabashnik, 2008). These

same issues will have to be addressed with transgenic citrus.

While scientists continuously seek better gene targets for suppression, preferably

targets that will elicit a desired response (reduced fitness, low fecundity, mortality) with

a low dose of dsRNA that is not injurious to non-target organisms, the cost of delivery of

these dsRNA molecules will be important. The persistence of the dsRNA molecules in

the plants and/or the environment will also be important in the choice of a delivery

strategy. The application of attenuated viral agents, like CTV that has the dsRNA

replicative form, implies a regular and cost-effective supply of dsRNA sequences will be

available, provided that the viral genomes do not lose the construct and revert to the

wild type (Burand and Hunter, 2013). The alternative of applying dsRNA in insecticidal

sprays is feasible, but the cost of producing large quantities of dsRNA on a commercial

basis is currently a deterrent.

Safety of the transgenes in my experiments will need to be empirically

determined. However, transgenic citrus rootstocks engineered for resistance to D.

abbreviatus and grafted to a non-transgenic fruit-producing scion will bear fruits without

transgenes. Therefore, evidence that recombinant products produced in the roots but

not mobilized through phloem or xylem will boost consumer acceptance. Cathepsins are

ubiquitous enzymes in numerous organisms (Beton et al., 2012), and therefore it is

possible that peptide inhibitors may also affect similar enzymes in non-target organisms

that feed on citrus. However, many similar peptide inhibitors, for example, legume

inhibitors (Ryan, 1973; Mithöfer and Boland, 2012), are present in foods that are safe

for consumption. The more prudent issue will be the levels of accumulation of peptide

110

inhibitors in edible tissues of transgenic citrus. The proposed vector for expression of D.

abbreviatus toxins in citrus, CTV, has co-evolved with citrus (Moreno et al., 2008) and

the distribution of the virus is throughout the plant (Dawson et al., 2013). Fuchs and

Gonsalves (2007) reported the use of mild strains of CTV for cross-protection of citrus in

Brazil for years. This and other virus-based plant protection practices, including

engineered viruses, are commonplace and no known adverse effects on human health

or environmental safety are reported.

The greatest obstacle to the sustainability of a transgenic citrus tactic could be

the development of resistance in populations of D. abbreviatus exposed to the

transgenic variety. Insects are known to be very adept at developing resistance. The

onset of resistance can in the least be delayed with the incorporation of multiple

insecticidal toxins with different modes of action in the transgenic citrus. In cross-

resistance, tolerance to an insecticidal toxin is enough to protect an individual from a

different, but similar acting toxin. Studies of a strain of the cotton bollworm, Helicoverpa

zea, revealed that resistance to the Bt toxin Cry1Ac protein had increased survivorship

of the pest on pyramided cotton plants as a result of cross resistance between the

Cry1A and another Bt toxin, Cry2A (Brévault et al., 2013). Cross-resistance will be

unlikely in the proposed three-pronged transgenic approach in citrus.

Finally, IPM strategies can prolong the sustainability of a transgenic citrus

strategy (Kos et al., 2009). The inclusion of refuges is acclaimed to have helped delay

insect resistance to Bt crops (Tabashnik et al., 2008). Cultivation of citrus using

transgenic strains may require minimal use of insecticides. However, minimal

insecticide use can function in tandem with biological control. Parasitoids and predators

111

would benefit from the reduced use of broad-spectrum insecticides used in transgenic

citrus groves and augment the control of the weevil. Transgenic citrus will therefore be

“IPM-friendly” at the same time as playing a pivotal role in suppressing pest populations

of D. abbreviatus. My studies bring the strategy of controlling D. abbreviatus with a

transgenic technique a step closer.

112

LIST OF REFERENCES

Abbott, W.S., 1925. A method of computing the effectiveness of an insecticide. J. Econ. Entomol. 18, 265-267.

Abdeen, A., Virgós, A., Olivella, E., Villanueva, J., Avilés, X., Gabarra, R., Prat, S., 2005. Multiple insect resistance in transgenic tomato plants over-expressing two families of plant proteinase inhibitors. Plant Mol. Biol. 57, 189–202.

Bachman, Bolognesi, R., Moar, W.J., Mueller, G.M., Paradise, M.S., Ramaseshadri, P., Tan, J., Uffman, J.P., Warren, J., Wiggins, B.E., Levine, S.L., 2013. Characterization of the spectrum of insecticidal activity of a double-stranded RNA with targeted activity against Western Corn Rootworm (Diabrotica virgifera virgifera LeConte). Transgenic Res. DOI 10.1007/s11248-013-9716-5.

Barrett, A.J., Kembhavi, A.A., Brown, M.A., Kirschke, H., Knight, C.G., Tamait, M., Hanadat, K., 1982. Analogues as inhibitors of cysteine proteinases including cathepsins B , H and L. Biochem. J. 201, 189–198.

Bates, S.L., Zhao, J.-Z., Roush, R.T., Shelton, A.M., 2005. Insect resistance management in GM crops: past, present and future. Nat. Biotechnol. 23, 57–62.

Baum, J. A., Bogaert, T., Clinton, W., Heck, G.R., Feldmann, P., Ilagan, O., Johnson, S., Plaetinck, G., Munyikwa, T., Pleau, M., Vaughn, T., Roberts, J., 2007. Control of coleopteran insect pests through RNA interference. Nat. Biotechnol. 25, 1322–1326.

Bautista, M.A.M., Miyata, T., Miura, K., Tanaka, T., 2009. RNA interference-mediated knockdown of a cytochrome P450, CYP6BG1, from the diamondback moth, Plutella xylostella, reduces larval resistance to permethrin. Insect Biochem. Mol. Biol. 39, 38–46.

Beavers, J.B., Selhime, A.G., 1975. Development of Diaprepes abbreviatus on potted citrus seedlings. Fla. Entomol. 58, 271–273.

Berger, A., Schechter, I., 1970. Mapping the active site of papain with the aid of peptide substrates and inhibitors. Philoso. Trans. R. Soc., Biol. Sci. 257, 249–264.

Beton, D., Guzzo, C.R., Ribeiro, A.F., Farah, C.S., Terra, W.R., 2012. The 3D structure and function of digestive cathepsin L-like proteinases of Tenebrio molitor larval midgut. Insect Biochem. Mol. Biol. 42, 655–664.

Boava, L.P., Cristofani-Yaly, M., Mafra, V.S., Kubo, K., Kishi, L.T., Takita, M. A., Ribeiro-Alves, M., Machado, M. A., 2011. Global gene expression of Poncirus trifoliata, Citrus sunki and their hybrids under infection of Phytophthora parasitica. BMC Genomics 12, 39.

113

Bogyo, M., Verhelst, S., Bellingard-Dubouchaud, V., Toba, S., Greenbaum, D., 2000. Selective targeting of lysosomal cysteine proteases with radiolabeled electrophilic substrate analogs. Chem. Biol. 7, 27–38.

Borovsky, D., Carlson, D.A., Griffin, R.R., Shabanowitz, J., Hunt, D.F., 1993. Mass spectrometry and characterization of Aedes aegypti trypsin modulating oostatic factor (TMOF) and its analogs. Insect Biochem. Mol. Biol. 23, 703-712.

Borovsky, D., Mahmood F., 1995. Feeding the mosquito Aedes aegypti with TMOF and its analogs ; effect on trypsin biosynthesis and egg development. Regul. Pept. 57, 273–281.

Borovsky, D., Meola, S.M., 2004. Biochemical and cytoimmunological evidence for the control of Aedes aegypti larval trypsin with Aea-TMOF. Arch. Insect Biochem. Physiol. 55, 124–139.

Borovsky, D., Janssen, I., Vanden Broeck, J., Huybrechts, R., Verhaert, P., De Bondt, H.L., Bylemans, D., De Loof, A., 1996. Molecular sequencing and modeling of Neobelleria bullata trypsin. Evidence for translational control by Neobellieria trypsin-modulating oostatic factor. Eur. J. Biochem. / FEBS 237, 279-287.

Bown, D.P., Wilkinson, H.S., Jongsma, M. A., Gatehouse, J. A., 2004. Characterisation of cysteine proteinases responsible for digestive proteolysis in guts of larval western corn rootworm (Diabrotica virgifera) by expression in the yeast Pichia pastoris. Insect Biochem. Mol. Biol. 34, 305–320.

Bradford, K.J., Van Deynze, A., Gutterson, N., Parrott, W., Strauss, S.H., 2005. Regulating transgenic crops sensibly: lessons from plant breeding, biotechnology and genomics. Nat. Biotechnol. 23, 439-455.

Bravo, A., Gill, S.S., Soberón, M., 2007. Mode of action of Bacillus thuringiensis Cry and Cyt toxins and their potential for insect control. Toxicon 49, 423-435.

Bravo, A., Likitvivatanavong, S., Gill, S.S., Soberón, M., 2011. Bacillus thuringiensis: A story of a successful bioinsecticide. Insect Biochem. Mol. Biol. 41, 423–431.

Brévault, T., Heuberger, S., Zhang, M., Ellers-kirk, C., Ni, X., Masson, L., Li, X., 2013. Potential shortfall of pyramided transgenic cotton for insect resistance management. Proc. Natl. Acad. Sci. U.S.A. 110, 5806-5811.

Burand, J.P., Hunter, W.B., 2013. RNAi: future in insect management. J. Invertebr. Pathol. 112, 68–74.

Butt, T.R., Edavettal, S.C., Hall, J.P., Mattern, M.R., 2005. SUMO fusion technology for difficult-to-express proteins. Protein Expr. Purif. 43, 1–9.

114

Carmona, E., Dufour, E., Plouffe, C., Takebe, S., Mason, P., Mort, J.S., Ménard, R., 1996. Potency and selectivity of the cathepsin L propeptide as an inhibitor of cysteine proteases. Biochem. 35, 8149–8157.

Choe, Y., Leonetti, F., Greenbaum, D.C., Lecaille, F., Bogyo, M., Brömme, D., Ellman, J. A, Craik, C.S., 2006. Substrate profiling of cysteine proteases using a combinatorial peptide library identifies functionally unique specificities. J. Biol. Chem. 281, 12824–12832.

Chougule, N.P., Li, H., Liu, S., Linz, L.B., Narva, K.E., Meade, T., Bonning, B.C., 2013. Retargeting of the Bacillus thuringiensis toxin Cyt2Aa against hemipteran insect pests. Proc. Natl. Acad. Sci. U.S.A. 110, 8465–8470.

Christou, P., Capell, T., Kohli, A., Gatehouse, J. A, Gatehouse, A.M.R., 2006. Recent developments and future prospects in insect pest control in transgenic crops. Trends Plant Sci. 11, 302–308.

Clara, R.O., Soares, T.S., Torquato, R.J.S., Lima, C. a, Watanabe, R.O.M., Barros, N.M.T., Carmona, A.K., Masuda, A., Vaz, I.S., Tanaka, A.S., 2011. Boophilus microplus cathepsin L-like (BmCL1) cysteine protease: specificity study using a peptide phage display library. Vet. Parasitol. 181, 291–300.

Clough, S.J., Bent, A. F., 1998. Floral dip: a simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J. 16, 735–43.

Cotrin, S.S., Puzer, L., de Souza Judice, W.A., Juliano, L., Carmona, A.K., Juliano, M.A., 2004. Positional-scanning combinatorial libraries of fluorescence resonance energy transfer peptides to define substrate specificity of carboxydipeptidases: assays with human cathepsin B. Anal. Biochem. 335, 244–252.

Coulombe, R., Li, Y., Takebe, S., Ménard, R., Mason, P., Mort, J.S., Cygler, M., 1996. Crystallization and preliminary X-ray diffraction studies of human procathepsin L. Proteins 25, 398-400.

Coutinho-Abreu, I.V., Sharma, N.K., Robles-Murguia, M., Ramalho-Ortigao, M., 2013. Characterization of Phlebotomus papatasi peritrophins, and the role of PpPer1 in Leishmania major survival in its natural vector. PLoS Negl. Trop. Dis. 7, e2132.

Coutinho-Abreu, I.V., Zhu, K.Y., Romalho-Ortiago, M., 2010. Transgenesis and paratransgenesis to control insect-borne diseases: Current status and future challenges. Parasitol. Int. 59, 1-8.

Crickmore, N., 2006. Beyond the spore – past and future developments of Bacillus thuringiensis as a biopesticide. J. Appl. Microbiol. 101, 616-619.

Cristofoletti, P.T., Ribeiro, A.F., Terra, W.R., 2005. The cathepsin L-like proteinases from the midgut of Tenebrio molitor larvae: sequence, properties,

115

immunocytochemical localization and function. Insect Biochem. Mol. Biol. 35, 883–901.

Dawson, W.O., Folimonova, S.Y., Folimonov, A.S., 2010. Viral-based transient-expression vector system for trees. United States Patent Application Publication, pp. 1-15.

Dawson, W.O., Garnsey, S.M., Tatineni, S., Folimonova, S.Y., Harper, S.J., Gowda, S., 2013. Citrus tristeza virus-host interactions. Front. Microbiol. 4, 1-10.

De Maagd, R. A., Bravo, A., Crickmore, N., 2001. How Bacillus thuringiensis has evolved specific toxins to colonize the insect world. Trends Genet. 17, 193–199.

Deraison, C., Darboux, I., Duportets, L., Gorojankina, T., Rahbé, Y., Jouanin, L., 2004. Cloning and characterization of a gut-specific cathepsin L from the aphid Aphis gossypii. Insect Mol. Biol. 13, 165–177.

Deshapriya,M.C., Takeuchi, A., Shirao, K., Isa, K., Watabe, S., Murakami, R., Tsujimura, H., Yamamoto, Y., 2007. Drosophila CTLA-2-like protein (D/CTLA-2) inhibits cysteine proteinase 1 (CP1), a cathepsin L-like enzyme. Zool. Sci. 24, 21-30.

Diaz, A.P., Mannion, C., Schaffer, B., 2006. Effect of root feeding by Diaprepes abbreviatus (Coleoptera : Curculionidae) larvae on leaf gas exchange and growth of three ornamental tree species. J. Econ. Entomol. 99, 811-821.

DiFazio, S.P., Leonardi, S., Slavov, G.T., Garman, S.L., Adams, W.T., Strauss, S.H., 2012. Gene flow and simulation of transgene dispersal from hybrid poplar plantations. N. Phytol. 193, 903–15.

Duan, G., Saint, R.B., Helliwell, C. A., Behm, C. A, Wang, M.-B., Waterhouse, P.M., Gordon, K.H.J., 2013. C. elegans RNA-dependent RNA polymerases rrf-1 and ego-1 silence Drosophila transgenes by differing mechanisms. Cell. Mol. Life Sci.  70, 1469–1481.

Edwards, M.G., Gatehouse, A.J., Gatehouse, A.M.R., 2010. Molecular and biochemical characterisation of a dual proteolytic system in vine weevil larvae (Otiorhynchus sulcatus Coleoptera: Curculionidae). Insect Biochem. Mol. Biol. 40, 785-791.

Ernst, J. A., Ascunce, M.S., Clark, A. M., Nigg, H.N., 2006. Polymorphic microsatellite loci for Diaprepes root weevil (Diaprepes abbreviatus L.). Mol. Ecol. Notes 6, 1–3.

Fire, A., Xu, S., Montgomery, M.K., Kostas, S.A., Driver, S.F., Mello, C.C., 1998. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nat. 391, 806-811.

116

Folimonov, A.S., Folimonova, S.Y., Bar-Joseph, M., Dawson, W.O., 2007. A stable RNA virus-based vector for citrus trees. Virol. 368, 205–216.

Fuchs, M., Gonsalves, D., 2007. Safety of virus-resistant transgenic plants two decades after their introduction: lessons from realistic field risk assessment studies. Annu. Rev. Phytopathol. 45, 173–202.

Gassmann, A.J., Carrière, Y., Tabashnik, B.E., 2009. Fitness costs of insect resistance to Bacillus thuringiensis. Annu. Rev. Entomol. 54, 147–163.

Gassmann, A.J., Petzold-Maxwell, J.L., Keweshan, R.S., Dunbar, M.W., 2011. Field-evolved resistance to Bt maize by western corn rootworm. PloS One 6, e22629.

Gatehouse, J. A., 2008. Biotechnological prospects for engineering insect-resistant plants. Plant Physiol.146, 881–887.

Georgis, R., Koppenhöfer, A. M., Lacey, L. A., Bélair, G., Duncan, L.W., Grewal, P.S., Samish, M., Tan, L., Torr, P., van Tol, R.W.H.M., 2006. Successes and failures in the use of parasitic nematodes for pest control. Biol. Control 38, 103–123.

Goodman, R.E., Vieths, S., Sampson, H. A., Hill, D., Ebisawa, M., Taylor, S.L., van Ree, R., 2008. Allergenicity assessment of genetically modified crops-what makes sense? Nat. Biotechnol. 26, 73–81.

Gordon, K.H.J., Waterhouse, P.M., 2007. RNAi for insect-proof plants. Nat. Biotechnol. 25, 1231–1232.

Graham, J.H., Bright, D.B., Pathology, P., McCoy, C.W., Alfred, L., 2003. Phytophthora – Diaprepes weevil complex : Phytophthora spp. relationship with citrus rootstocks. Plant Dis. 87, 85-90.

Gewies, A., Grimm, S., 2003. Cathepsin-B and cathepsin-L expression levels do not correlate with sensitivity of tumour cells to TNF-α-mediated apoptosis. Br. J. Cancer 89, 1574-1580.

Hall, D.G., 1995. A revision to the bibliography of the sugarcane rootstalk borer weevil, Diaprepes abbreviatus (Coleoptera : Curculionidae). Fla. Entomol. 78, 364–377.

Hall, D.G., Peña, J., Franqui, R., Nguyen, R., Stansly, P., 2001. Status of Biol. Control by egg parasitoids of Diaprepes abbreviatus (Coleoptera : Curculionidae) in citrus in Fla. and Puerto Rico. BioControl 46, 61–70.

Haq, S.K., Atif, S.M., Khan, R.H., 2004. Protein proteinase inhibitor genes in combat against insects, pests, and pathogens: natural and engineered phytoprotection. Arch. Biochem. Biophys. 431, 145–159.

Harlander, S.K., 2002. The evolution of modern agriculture and its future with biotechnology. J. Am. Coll. Nutr. 21, 161–165.

117

Hasnain, S., Hirama, T., Tam, A., Mort, J.S., 1992. Characterization of recombinant rat cathepsin B and nonglycosylated mutants expressed in yeast. J. Biol. Chem. 267, 4713-4721.

Hunt, I., 2005. From gene to protein: a review of new and enabling technologies for multi-parallel protein expression. Protein Expr. Purif. 40, 1–22.

Huvenne, H., Smagghe, G., 2010. Mechanisms of dsRNA uptake in insects and potential of RNAi for pest control: a review. J. Insect Physiol. 56, 227–235.

James, C., 2012. Global status of commercialized biotech / GM crops. ISAAA Briefs 44, pp. 1-18.

Jetter, K.M., Godfrey, K., 2009. Diaprepes root weevil, a new California pest, will raise costs for pest control and trigger quarantines. Calif. Agric. 63, 121–126.

Kamboj, R.C., Pal, S., Raghav, N., Singh, H., 1993. A selective colorimetric assay for cathepsin L using Z-Phe-Arg-4-methoxy-beta-naphthylamide. Biochim. 75, 873–878.

Kapust, R.B., Waugh, D.S., 1999. Escherichia coli maltose-binding protein is uncommonly effective at promoting the solubility of polypeptides to which it is fused. Protein Sci.  8, 1668–1674.

Kathage, J., Qaim, M., 2012. Economic impacts and impact dynamics of Bt (Bacillus thuringiensis) cotton in India. Proc. Natl. Acad. Sci. U.S.A. 109, 11652–11656.

Kaulmann, G., Palm, G.J., Schilling, K., Hilgenfeld, R., Wiederanders, B., 2006. The crystal structure of a Cys25→Ala mutant of human procathepsin S elucidates enzyme-prosequence interactions. Protein Sci. 15, 2619-2629.

Keene, K.M., Foy, B.D., Sanchez-Vargas, I., Beaty, B.J., Blair, C.D., Olson, K.E., 2004. RNA interference acts as a natural antiviral response to O’nyong-nyong virus (Alphavirus; Togaviridae) infection of Anopheles gambiae. Proc. Natl. Acad. Sci. U.S.A. 101, 17240–17245.

Kirschke, H., Langner, J., Wiederanders, B., Ansorge, S., Bohley, P., 1977. Cathepsin L: a new proteinase from rat liver lysosomes. Eur. J. Biochem. 74, 293-301.

Koiwa, H., Shade, R.E., Zhu-Salzman, K., D’Urzo, M.P., Murdock, L.L., Bressan, R. A., Hasegawa, P.M., 2000. A plant defensive cystatin (soyacystatin) targets cathepsin L-like digestive cysteine proteinases (DvCALs) in the larval midgut of western corn rootworm (Diabrotica virgifera virgifera). FEBS Lett. 471, 67–70.

Kos, M., van Loon, J.J.A., Dicke, M., Vet, L.E.M., 2009. Transgenic plants as vital components of integrated pest management. Trends Biotechnol. 27, 621–627.

118

Knorr, E., Bingsohn, L., Kanost, M.R., Vilcinskas, A., 2013. Tribolium castaneum as a Model for High-Throughput RNAi Screening. Adv. Biochem. Eng. Biotechnol. doi:10.1007/10.

Krieger, E., Koraimann, G., Vriend, G., 2002. Increasing the precision of comparative models with YASARA NOVA - a self-parameterizing force field. Proteins 47, 393-402.

Kulkarni, M.M., Booker, M., SIlver, S.J., Friedman, A., Hong, P., Perrimon, N., Mathey-Prerot, B. 2006. Evidence of off-target effects associated with long dsRNAs in Drosophila melanogaster cell-based assays. Nat. Methods 3, 833-838.

Kurata, M., Yamamoto, Y., Watabe, S., Makino, Y., Ogawa, K., Takahashi, S.Y. 2001. Bombyx cysteine proteinase inhibitor (BCPI) homologous to propeptide regions of cysteine proteinases is a strong, selective inhibitor of cathepsin L-like cysteine proteinases. J. Biochem. 130, 857-863.

Lacey, L.A., Frutos, R., Kaya, H.K., Vail, P., 2001. Insect pathogens as biological control agents: do they have a future? Biol. Control 21, 230-248.

Lapointe, S.L., 2000. Particle film deters oviposition by Diaprepes abbreviatus (Coleoptera : Curculionidae). J. Econ. Entomol. 93, 1459–1463.

Lapointe, S.L., Borchert, D.M., Hall, D.G., 2007. Effect of low temperatures on mortality and oviposition in conjunction with climate mapping to predict spread of the root weevil Diaprepes abbreviatus and introduced natural enemies. Environ. Entomol. 36, 73-82.

Lapointe, S.L., Bowman, K.D., 2002. Is there meaningful plant resistance to Diaprepes abbreviatus (Coleptera: Curculionidae) in citrus rootstock germplasm? J. Econ. Entomol. 95, 1059-1065.

Lapointe, S.L., Shapiro, J.P., 1999. Effect of Soil Moisture on Development of Diaprepes abbreviatus (Coleoptera: Curculionidae). Fla. Entomol. 82, 291-299.

Larson, E.T., Parussini, F., Huynh, M.H., Giebel, J.D., Kelley, A.M., Zhang, L., Bogyo, M., Merritt, E.A., Carruthers, V.B., 2009. Toxoplasma gondii cathepsin L is the primary target of the invasion-inhibitory compound morpholinurea-leucyl-homophenyl-vinyl sulfone phenyl. J. Biol. Chem. 284, 26839-26850.

Laskowski R.A., MacArthur M.W., Moss D.S., Thornton J.M., 1993. PROCHECK: a program to check the stereochemistry of protein structures. J. Appl. Crystallogr. 26, 283-291.

Liu, F., Xu, Z., Zhu, Y.C., Huang, F., Wang, Y., Li, Huiling, Li, Hua, Gao, C., Zhou, W., Shen, J., 2010. Evidence of field-evolved resistance to Cry1Ac-expressing Bt cotton in Helicoverpa armigera (Lepidoptera: Noctuidae) in northern China. Pest Manag. Sci. 66, 155–161.

119

Lu, Y., Wu, K., Jiang, Y., Guo, Y., Desneux, N., 2012. Widespread adoption of Bt cotton and insecticide decrease promotes biocontrol services. Nat. 487, 362–365.

Macedo, M.R.L., Freire, M.D.G.M., Cabrini, E.C., Toyama, M.H., Novello, J.C., Marangoni, S., 2003. A trypsin inhibitor from Peltophorum dubium seeds active against pest proteases and its effect on the survival of Anagasta kuehniella (Lepidoptera: Pyralidae). Biochim. Biophys. Acta 1621, 170–182.

Manceva, S.D., Pusztai-Carey, M., Russo, P.S., Butko, P., 2005. A detergent-like mechanism of action of the cytolytic toxin Cyt1A from Bacillus thuringiensis var. israelensis. Biochem. 589–597.

Mao, Y.-B., Cai, W.-J., Wang, J.-W., Hong, G.-J., Tao, X.-Y., Wang, L.-J., Huang, Y.-P., Chen, X.-Y., 2007. Silencing a cotton bollworm P450 monooxygenase gene by plant-mediated RNAi impairs larval tolerance of gossypol. Nat. Biotechnol. 25, 1307–1313.

Marroquin, L.D., Elyassnia, D., Griffitts, J.S., Feitelson, J.S., Aroian, R. V, 2000. Bacillus thuringiensis (Bt) toxin susceptibility and isolation of resistance mutants in the nematode Caenorhabditis elegans. Genet. 1699, 1693–1699.

Martin, C.G., Mannion, C., Schaffer, B., 2009. Effects of herbivory by Diaprepes abbreviatus (Coleoptera: Curculionidae) larvae on four woody ornamental plant species. J. Econ. Entomol. 102, 1141–1150.

Meister, G., Tuschl, T., 2004. Mechanisms of gene silencing by double-stranded RNA. Nat. 431, 343–349.

Melo, F., Feytmans, E., 1998. Assessing protein structures with a non-local atomic interaction energy. J. Mol. Biol. 277, 1141-1152.

Mithöfer, A., Boland, W., 2012. Plant defense against herbivores: chemical aspects. Annu. Rev. Plant Biol. 63, 431–450.

Miyaji, T., Murayama S., Kouzuma Y., Kimura, N., Kanost. M.R., Kramer, K.J., Yonekura, M. 2010. Molecular cloning of a multidomain cysteine protease and protease inhibitor precursor gene from the tobacco hornworm (Manduca sexta) and functional expression of the cathepsin F-like cysteine protease domain. Insect Biochem. Mol. Biol. 40, 835-846.

Moreno, P., Ambros, S., Albiach-Marti, M.R., Guerri, J., Peña, L., 2008. Citrus tristeza virus : a pathogen that changed the course of the citrus industry. Mol. Plant Pathol. 9, 251–268.

Murdock, L.L., Shade, R.E., 2002. Lectins and protease inhibitors as plant defenses against insects. J. Agric. Food Chem. 50, 6605-6611.

120

Nigg, H.N., Siméson, S.E., Ramos, L.E., Tomerlin, A.T., Cuyler, N.W., Alfred, L., 1999. Fipronil for Diaprepes abbreviatus (Coleoptera: Curculionidae ) larval control in container-grown citrus. Proc. Fla. State Hortic. Soc. 112, 77-79.

Nunes, F., Aleixo, A., Barchuk, A., Bomtorin, A., Grozinger, C., Simões, Z., 2013. Non-target effects of green fluorescent protein (GFP)-derived double-stranded RNA (dsRNA-GFP) used in honey bee RNA interference (RNAi) assays. Insects 4, 90–103.

Obbard, D.J., Gordon, K.H.J., Buck, A.H., Jiggins, F.M., 2009. The evolution of RNAi as a defence against viruses and transposable elements. Phil. Trans. R. Soc. B 364, 99–115.

Oppert, B., Morgan, T.D., Culbertson, C., Kramer, K.J., 1993. Dietary mixtures of cysteine and serine proteinase inhibitors exhibit synergistic toxicity toward the red flour beetle, Tribolium castaneum. Comp. Biochem. Physiol. 105, 379–385.

Otto, H.-H., Schirmeister, T., 1997. Cysteine proteases and their inhibitors. Chem. Rev. 97, 133–172.

Pettersen E.F., Goddard T.D., Huang C.C., Couch G.S., Greenblatt D.M., Meng E.C., Ferrin T.E., 2004. UCSF Chimera - a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605-1612.

Pérez, C., Fernandez, L.E., Sun, J., Folch, J.L., Gill, S.S., Soberón, M., Bravo, A., 2005. Bacillus thuringiensis subsp. israelensis Cyt1Aa synergizes Cry11Aa toxin by functioning as a membrane-bound receptor. Proc. Natl. Acad. Sci. U.S.A. 102, 18303-18308.

Philip, J.M.D., Fitches, E., Harrison, R.L., Bonning, B., Gatehouse, J.A., 2007. Characterisation of functional and insecticidal properties of a recombinant cathepsin L-like proteinase from flesh fly (Sarcophaga peregrina), which plays a role in differentiation of imaginal discs. Insect Mol. Biol. 37, 589-600.

Price, D.R.G., Gatehouse, J. A., 2008. RNAi-mediated crop protection against insects. Trends Biotechnol. 26, 393–400.

Pyati, P.S., Bell, H.A., Fitches, E., Price, D.R.G., Gatehouse, D.R.G., Gatehouse, J.A., 2009. Cathepsin L-like cysteine proteinase (DcCathL) from Delia coarctata (wheat bulb fly): Basis of insecticidal activity. Insect Biochem. Mol. Biol. 39, 535-546.

Qiu, S., Adema, C.M., Lane, T., 2005. A computational study of off-target effects of RNA interference. Nucleic Acids Res. 33, 1834–1847.

Rao, R., Fiandra, L., Giordana, B., de Eguileor, M., Congiu, T., Burlini, N., Arciello, S., Corrado, G., Pennacchio, F., 2004. AcMNPV ChiA protein disrupts the

121

peritrophic membrane and alters midgut physiology of Bombyx mori larvae. Insect Biochem. Mol. Biol. 34, 1205–1213.

Rawlings, N.D., Barrett, A.J., Bateman, A., 2012. MEROPS: The database of proteolytic enzymes, their substrates and inhibitors. Nucleic Acids Res. 40, 343–350.

Romeis, J., McLean, M.A., Shelton, A.M., 2013. When bad science makes good headlines: Bt maize and regulatory bans. Nat. Biotechnol. 31, 386–387.

Rommens, C.M., 2010. Barriers and paths to market for genetically engineered crops. Plant Biotechnol. J. 8, 101–111.

Roush, R.T., 1997. Bt-transgenic crops: just another pretty insecticide or a chance for a new start in resistance management? Pestic. Sci. 51, 328-334.

Rozman-Pungercar, J., Kopitar-Jerala, N., Bogyo, M., Turk, D., Vasiljeva, O., Stefe, I., Vandenabeele, P., Brömme, D., Puizdar, V., Fonović, M., Trstenjak-Prebanda, M., Dolenc, I., Turk, V., Turk, B., 2003. Inhibition of papain-like cysteine proteases and legumain by caspase-specific inhibitors: when reaction mechanism is more important than specificity. Cell Death Differ. 10, 881–888.

Ryan, C.A., 1990. Protease inhibitors in plants: Genes for improving defenses against insects and pathogens. Annu. Rev. Phytopathol. 28, 425–449.

Schlüter, U., Benchabane, M., Munger, A., Kiggundu, A., Vorster, J., Goulet, M.-C., Cloutier, C., Michaud, D., 2010. Recombinant protease inhibitors for herbivore pest control: a multitrophic perspective. J. Exp. Bot. 61, 4169–4183.

Sciabola, S., Cao, Q., Orozco, M., Faustino, I., Stanton, R. V., 2013. Improved nucleic acid descriptors for siRNA efficacy prediction. Nucleic Acids Res. 41, 1383–94.

Shabalina, S. A., Koonin, E. V., 2008. Origins and evolution of eukaryotic RNA interference. Trends Ecol. Evol. 23, 578–587.

Shapiro, J.P., Gottwald, T.R., 1995. Citrus cultivars resistance rootstock. J. Econ. Entomol. 8, 148–154.

Shapiro-Ilan, D.I., Gouge, D.H., Piggott, S.J., Fife, J.P., 2006. Application technology and environmental considerations for use of entomopathogenic nematodes in biological control. Biol. Control. 38, 124-133.

Shaw, E., Mohanty, S., Colic, A., Stoka, V., Turk, V., 1993. The affinity-labelling of cathepsin S with peptidyl diazomethyl ketones. Comparison with the inhibition of cathepsin L and calpain. FEBS Lett. 334, 340–342.

Shelton, A.M., Zhao, J-Z., Roush, R.T. 2002. Economic, ecological, food safety, and social consequences of the deployment of Bt transgenic plants. Annu. Rev. Evol. 47, 845-881.

122

Sim, C., Denlinger, D.L., 2008. Insulin signaling and FOXO regulate the overwintering diapause of the mosquito Culex pipiens. Proc. Natl. Acad. Sci. U.S.A. 105, 6777–6781.

Sirot, L.K., Lapointe, S.L., 2008. Patterns and consequences of mating behavior of the root weevil Diaprepes abbreviatus (Coleoptera: Curculionidae). Fla. Entomol. 91, 400-407.

Simpson, S.E., Nigg, H.N., Coile, N.C., Adair, R.C., 1996. Diaprepes abbreviatus (Coleoptera: Curculionidae): host plant associations. Environ. Entomol. 25, 333–349.

Soberón, M., López-Díaz, J. a, Bravo, A., 2013. Cyt toxins produced by Bacillus thuringiensis: a protein fold conserved in several pathogenic microorganisms. Pept. 41, 87–93.

Stachowiak, K., Tokmina, M., Karpińska, A., Sosnowska, R., Wiczk, W., 2004. Fluorogenic peptide substrates for carboxydipeptidase activity of cathepsin B. Acta Biochim. Pol. 51, 81–92.

Stack, C.M., Caffrey, C.R., Donnelly, S.M., Seshaadri, A., Lowther, J., Tort, J.F., Collins, P.R., Robinson, M.W., Xu, W., McKerrow, J.H., Craik, C.S., Geiger, S.R., Marion, R., Brinen, L.S., Dalton, J.P., 2008. Structural and functional relationships in the virulence-associated cathepsin L proteases of the parasitic liver fluke Fasciola hepatica. J. Biol. Chem. 283, 9896-9908.

Stuart, R.J., McCoy, C.W., Castle, W.S., Graham, J.H., Rogers, M.E., 2006. Diaprepes, Phytophthora, and hurricanes: rootstock selection and pesticide use affect growth and survival of 'hamlin' orange trees in a Central Florida. citrus grove. Proc. Fla. State Hortic. Soc. 119, 128-135.

Stuart, R.J., Shapiro-Ilan, D.I., James, R.R., Nguyen, K.B., McCoy, C.W., 2004. Virulence of new and mixed strains of the entomopathogenic nematode Steinernema riobrave to larvae of the citrus root weevil Diaprepes abbreviatus. Biol. Control 30, 439–445.

Tabashnik, B.E., 2008. Delaying insect resistance to transgenic crops. Proc. Natl. Acad. Sci. U.S.A. 105, 19029–19030.

Tabashnik, B.E., Gassmann, A.J., Crowder, D.W., Carriére, Y., 2008. Insect resistance to Bt crops: evidence versus theory. Nat. Biotechnol. 26, 199–202.

Tabashnik, B.E., Morin, S., Unnithan, G.C., Yelich, A.J., Ellers-Kirk, C., Harpold, V.S., Sisterson, M.S., Ellsworth, P.C., Dennehy, T.J., Antilla, L., Liesner, L., Whitlow, M., Staten, R.T., Fabrick, J.A., Li, X., Carrière, Y., 2012. Sustained susceptibility of pink bollworm to Bt cotton in the United States. GM Crop Food: Biotechnol. Agric. Food Chain 3, 194–200.

123

Tao, K., Stearns N.A., Dong, J.M., Wu, Q.I., Sahagian, G.G., 1994. The proregion of cathepsin L is required for proper folding, stability, and ER exit. Arch. Biochem. Biophys. 311, 19-27.

Tatineni, S., Gowda, S., Dawson, W.O., 2010. Heterologous minor coat proteins of citrus tristeza virus strains affect encapsidation, but the coexpression of HSP70h and p61 restores encapsidation to wild-type levels. Virol. 402, 262-270.

Tchoupé, J.R., Moreau, T., Gauthier, F., Bieth, J.G., 1991. Photometric or fluorometric assay of cathepsin B, L and H and papain using substrates with an aminotrifluoromethylcoumarin leaving group. Biochim. Biophys. Acta 1076, 149–151.

Terenius, O., Papanicolaou, A., Garbutt, J.S., Eleftherianos, I., Huvenne, H., Kanginakudru, S., Albrechtsen, M., An, C., Aymeric, J.-L., Barthel, A., Bebas, P., Bitra, K., Bravo, A., Chevalier, F., Collinge, D.P., Crava, C.M., de Maagd, R.A., Duvic, B., Erlandson, M., Faye, I., Felföldi, G., Fujiwara, H., Futahashi, R., Ghande, A.S., Gatehouse, H.S., Gatehouse, L.N., Giebultowicz, J.M., Goméz, I., Grimmelikhuijzen, C.J.P., Groot, A.T., Hauser, F., Heckel, D.J., Hegedus, D.D., Hrycaj, S., Huang, L., Hull, J.J., Iaotrou K., Iga, M., Kanost, M.R., Kotwica, J., Li, C., Li, J., Liu, J., Lundmark, M., Matsumoto, S., Meyering-Vos, M., Millichap, P.J., Montiero, A., Mrinal, N., Niimi, T., Nowara, D., Ohnishi A., Oostra, V., Ozaki, K., Papakonstantinou, M., Popadic, A., Rajam, M.V., Saenko, S., Simpson, R.M., Soberón, M., Strand, M., Tomita, M., Toprak, U., Wang, P., Wee, C.W., Whyard, S., Zhang, W., Nagaraju, J., ffrench-Constant, R.H., Herrero, S., Gordon, K., Swevers, L., Smagghe, G., 2011. RNA interference in Lepidoptera: an overview of successful and unsuccessful studies and implications for experimental design. J. Insect Physiol. 57, 231–45.

Terra, W.R., 1990. Evolution of digestive systems in insects. Ann. Rev. Entomol. 35, 181-200.

Terra, W.R., 2001. The origin and functions of the insect peritrophic membrane and peritrophic gel. Arch. Insect Biochem. Physiol. 47, 47–61.

Tomoyasu, Y., Denell, R.E., 2004. Larval RNAi in Tribolium (Coleoptera) for analyzing adult development. Dev. Genes Evol. 214, 575–578.

Tomoyasu, Y., Miller, S.C., Tomita, S., Schoppmeier, M., Grossmann, D., Bucher, G., 2008. Exploring systemic RNA interference in insects: a genome-wide survey for RNAi genes in Tribolium. Genome Biol. 9, R10.1-22.

Turk, B., Turk, D., Turk, V., 2000. Lysosomal cysteine proteases: more than scavengers. Biochim. Biophys. Acta 1477, 98–111.

Turk, V., Stoka, V., Vasiljeva, O., Renko, M., Sun, T., Turk, B., Turk, D., 2012. Cysteine cathepsins: from structure, function and regulation to new frontiers. Biochim. Biophys. Acta 1824, 68–88.

124

Ulmer B.J., Duncan, R., Pavis, C. and Pena, J.., 2006. Egg parasitoids of citrus weevils in Guadeloupe. Fla. Entomol. 92, 311–314.

Vandesompele, J., De Preter, K., Pattyn, F., Poppe, B., Van Roy, N., De Paepe, A., Speleman, F., 2002. Accurate normalization of real-time quantitative RT-PCR data by geometric averaging of multiple internal control genes. Genome Biol. 3, 1-11.

Vaughn, T., Cavato, T., Brar, G., Coombe, T., DeGooyer, T., Ford, S., Groth, M., Howe, A., Johnson, S., Kolacz, K., Pilcher, C., Purcell, J., Romano, C., English, L., Pershing, J., 2005. A method of controlling corn rootworm feeding using a protein expressed in transgenic maize. Crop Sci. 45, 931-938.

Vila, L., Quilis, J., Meynard, D., Breitler, J.C., Marfà, V., Murillo, I., Vassal, J.M., Messeguer, J., Guiderdoni, E., San Segundo, B., 2005. Expression of the maize proteinase inhibitor (mpi) gene in rice plants enhances resistance against the striped stem borer (Chilo suppressalis): effects on larval growth and insect gut proteinases. Plant Biotechnol. J. 3, 187–202.

Weathersbee, A.A. III, Lapointe, S.L., Shatters, R.G. Jr, 2006. Activity of Bacillus thuringiensis isolates against Diaprepes abbreviatus (Coleoptera : Curculionidae). Fla. Entomol. 89,441-448.

Weissling, T.J., Peña, J.E., Knapp, J.L., 2009. Diaprepes Root Weevil, Diaprepes abbreviatus (Linnaeus) (Insecta: Coleoptera: Curculionidae). UF-IFAS Extension, 1-5.

Whyard, S., Singh, S., Wong, S., 2009. Ingested dsRNAa can act as species-specific insecticides. Insect Biochem. Mol. Biol. 39, 824-832.

Xue, J.-L., Cai, Q.-X., Zheng, D.-S., Yuan, Z.-M., 2005. The synergistic activity between Cry1Aa and Cry1c from Bacillus thuringiensis against Spodoptera exigua and Helicoverpa armigera. Lett. Appl. Microbiol. 40, 460–465.

Yan, J., Yuan, F., Long, G., Qin, L., Deng, Z., 2012. Selection of reference genes for quantitative real-time RT-PCR analysis in citrus. Mol. Biol. Rep. 39, 1831-1838.

Yan, X.H., De Bondt, H.L., Powell, C.C., Bullock, R.C., Borovsky, D., 1999. Sequencing and characterization of the citrus weevil, Diaprepes abbreviatus, trypsin cDNA. Effect of Aedes trypsin modulating oostatic factor on trypsin biosynthesis. Eur. J. Biochem. / FEBS 262, 627–636.

Zhang, X., Zhang, J., Zhu, K.Y., 2010. Chitosan/double-stranded RNA nanoparticle-mediated RNA interference to silence chitin synthase genes through larval feeding in the African malaria mosquito (Anopheles gambiae). Insect Mol. Biol. 19, 683–693.

125

Zhou, X., Oi, F.M., Scharf, M.E., 2006. Social exploitation of hexamerin: RNAi reveals a major caste-regulatory factor in termites. Proc. Natl. Acad. Sci. U.S.A. 103, 4499–4504.

Zhou, X., Smith, J. a, Oi, F.M., Koehler, P.G., Bennett, G.W., Scharf, M.E., 2007. Correlation of cellulase gene expression and cellulolytic activity throughout the gut of the termite Reticulitermes flavipes. Gene 395, 29–39.

Zhou, Z., Pang, J., Guo, W., Zhong, N., Tian, Y., Xia, G., Wu, J., 2012. Evaluation of the resistance of transgenic potato plants expressing various levels of Cry3A against the Colorado potato beetle (Leptinotarsa decemlineata Say) in the laboratory and field. Pest Manag. Sci. 68, 1595–1604.

Zhu-Salzman, K., Koiwa, H., Salzman, R. A., Shade, R.E., Ahn, J.-E., 2003. Cowpea bruchid Callosobruchus maculatus uses a three-component strategy to overcome a plant defensive cysteine protease inhibitor. Insect Mol. Biol. 12, 135–45.

126

BIOGRAPHICAL SKETCH

Sulley K. Ben-Mahmoud was born in Accra, Ghana. In August 2001, he enrolled

at the University of Ghana, and received his B.Sc. in Biochemistry in June of 2005. He

worked as a teaching assistant at the Biochemistry Department for a year and enrolled

into the entomology program at the University of Ghana under the guidance of Dr.

W.S.K. Gbewonyo in August 2006. He graduated with an M.Phil. in Entomology in June

2009, and was admitted in the same year to the University of Florida, initially under the

mentorship of Dr. Dov Borovsky (retired) at the Florida Medical Entomology Laboratory

in Vero Beach, FL. He switched to the guidance of Dr. Ronald D. Cave in May of 2011

at the Indian River Research and Education Center in Fort Pierce, FL. He did the

majority of his research under the supervision of Dr. Robert G. Shatters Jr. at the United

States Department of Agriculture-Horticultural Research Laboratory, Fort Pierce, FL. He

received his Ph.D. from the University of Florida in the fall of 2013.