Advances in Insect Physiology Volume 47 Dhadialla, Tarlochan S. - Gill, Sarjeet S-Insect Midgut and Insecticidal Proteins,-Academic Press, Elsevier 2014
Advances in Insect Physiology Volume 47 Dhadialla, Tarlochan S. - Gill, Sarjeet S-Insect Midgut and Insecticidal Proteins,-Academic Press, Elsevier 2014
No part of this publication may be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopying, recording, or any information storage and
retrieval system, without permission in writing from the publisher. Details on how to seek
permission, further information about the Publishers permissions policies and our
arrangements with organizations such as the Copyright Clearance Center and the Copyright
Licensing Agency, can be found at our website: www.elsevier.com/permissions.
This book and the individual contributions contained in it are protected under copyright by
the Publisher (other than as may be noted herein).
Notices
Knowledge and best practice in this field are constantly changing. As new research and
experience broaden our understanding, changes in research methods, professional practices,
or medical treatment may become necessary.
Practitioners and researchers must always rely on their own experience and knowledge in
evaluating and using any information, methods, compounds, or experiments described
herein. In using such information or methods they should be mindful of their own safety and
the safety of others, including parties for whom they have a professional responsibility.
To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors,
assume any liability for any injury and/or damage to persons or property as a matter of
products liability, negligence or otherwise, or from any use or operation of any methods,
products, instructions, or ideas contained in the material herein.
ISBN: 978-0-12-800197-4
ISSN: 0065-2806
Michael J. Adang
Department of Entomology, and Department of Biochemistry and Molecular Biology,
University of Georgia, Athens, Georgia, USA
Md. Shohidul Alam
Division of Chemistry and Structural Biology, Institute for Molecular Bioscience,
The University of Queensland, St. Lucia, Queensland, Australia
James A. Baum
Monsanto Company, Chesterfield, Missouri, USA
Niraj S. Bende
Division of Chemistry and Structural Biology, Institute for Molecular Bioscience,
The University of Queensland, St. Lucia, Queensland, Australia
Colin Berry
Cardiff School of Biosciences, Cardiff University, Cardiff, United Kingdom
Neil Crickmore
School of Life Sciences, University of Sussex, Falmer, Brighton, United Kingdom
Rhoel R. Dinglasan
Department of Molecular Microbiology and Immunology, The Johns Hopkins Bloomberg
School of Public Health, Baltimore, Maryland, USA
Andrea J. Dowling
Biosciences, University of Exeter, Cornwall, United Kingdom
Richard H. ffrench-Constant
Biosciences, University of Exeter, Cornwall, United Kingdom
Volker Herzig
Division of Chemistry and Structural Biology, Institute for Molecular Bioscience,
The University of Queensland, St. Lucia, Queensland, Australia
Juan Luis Jurat-Fuentes
Department of Entomology and Plant Pathology, University of Tennessee, Knoxville,
Tennessee, USA
Robert M. Kennedy
Vestaron Corporation, Kalamazoo, Michigan, USA
Glenn F. King
Division of Chemistry and Structural Biology, Institute for Molecular Bioscience,
The University of Queensland, St. Lucia, Queensland, Australia
Paul J. Linser
The University of Florida Whitney Laboratory, St. Augustine, Florida, USA
vii
viii Contributors
Thomas Meade
Dow AgroSciences, LLC., Indianapolis, Indiana, USA
Kenneth E. Narva
Dow AgroSciences, LLC., Indianapolis, Indiana, USA
Leda Regis
Centro de Pesquisas Aggeu Magalhaes-Fiocruz, Recife-Pernambuco, Brazil
James K. Roberts
Monsanto Company, Chesterfield, Missouri, USA
Maria Helena Neves Lobo Silva Filha
Centro de Pesquisas Aggeu Magalhaes-Fiocruz, Recife-Pernambuco, Brazil
Nicholas P. Storer
Dow AgroSciences, LLC., Indianapolis, Indiana, USA
H. William Tedford
Vestaron Corporation, Kalamazoo, Michigan, USA
Yidong Wu
College of Plant Protection, Nanjing Agricultural University, Nanjing, China
PREFACE
The idea for this volume on Insect midgut and insecticidal proteins was
conceived from the realization that not a single source of reviews covers the
insect midgut and insecticidal proteins isolated from bacteria or arthropods.
This volume benefits anyone researching to find solutions for insect pest
control in agriculture and in public health.
The first chapter reviews Insect gut structure, function, development
and target of biological toxins. The insect midgut is the first barrier or a
target for ingested toxophores (small-molecule insecticides or insecticidal
proteins). For insecticidal proteins from the bacteria, Bacillus, Lysinibacillus
and Photorhabdus, the midgut provides several target sites by which these
proteins manifest their toxic action. However, for many target sites in other
tissues, the midgut can be a barrier for efficient delivery, like peptides from
spider venom (Chapter 8). Although a lot of the information reviewed
here is from mosquito and Drosophila midguts, these approaches and under-
standing can help draw parallels and differences between phytophagous
insects (agriculturally important) versus hematophagus insects (medical
importance). Additional proteomic studies on the midgut to identify and
characterize putative target sites would be beneficial for developing or
discovering alternate mechanisms of action. Linser and Dinglasan have pro-
vided an excellent review of the insect midgut with a discussion of possible
target sites.
Chapters 25 review various aspects of insecticidal proteins from Bacillus
and Lysinibacillus. In Chapter 2, Adang et al. review the diversity of insec-
ticidal proteins (three domain crystal (Cry), Cytolytic (Cyt), Binary Cry and
other parasporal toxins) from Bacillus. They review the mode of action of
these proteins, providing similarities and differences in the receptors used
for manifesting toxicity. The identification and characterization of toxin
receptors is important not only to create opportunities for discovering newer
toxins but also to modify known toxins to target insect pests that are less or
non-susceptible. Moreover, such investigations allow the development of
strategies to overcome or delay the development of resistance to insecticidal
proteins.
In Chapter 3, Filha et al. review the Binary (Bin) proteins from
Lysinibacillus sphaericus (Ls) that are mosquitocidal. The authors discuss the
structure, function and mechanisms by which these proteins cause toxicity
ix
x Preface
insecticides for interfering with the function of target site proteins. How-
ever, the use of dsRNAi provides a much higher level of selective toxicity
to insect pests and serves as an attractive approach. Although there are no
commercial products harnessing this approach as yet, it will not be long
before such products are commercially available.
In the last chapter (Chapter 6) related to Bt insecticidal proteins, Wu
reviews resistance development and resistance management strategies for
transgenic crops carrying Bt genes. The development of resistance is inev-
itable, and the challenge faced is how strategies can be deployed to delay the
targeted insect pests from developing resistance to the insecticidal proteins in
host transgenic crops. In this respect, it is also important to understand the
mechanisms and target site receptors/proteins these insect toxins use for
manifesting insecticidal activity and the mechanisms that lead to resistance
development. This aspect of resistance ties very well with the review in
Chapter 1 on mode of action of Bt proteins.
Chapters 7 and 8 review alternate sources of insecticidal proteins or pep-
tides. The discovery of insecticidal proteins from the bacteria Photorhabdus
and Xenorhabdus created a lot of interest among academic labs and industry
to understand the structurefunction and mode of action of these very large
(molecular size) and complex proteins. This is reviewed in Chapter 7.
Although genes encoding these proteins or their peptides have not been used
as transgenes in crops to control specific insect pests, the information gen-
erated can be leveraged with new approaches and capabilities to possibly
make use of such genes (modified or unmodified). ffrench-Constant and
Dowling have provided an extensive review of the many proteins from
the two bacteria, high-resolution structures and possible mechanisms of
action of the insecticidal proteins.
In Chapter 8 on Methods for deployment of spider venom peptides as
bioinsecticides, the authors describe a novel source of peptides from spider
venom that show very interesting and selective toxic activities in insects.
Most of these act on neuropeptide targets and provide a challenging oppor-
tunity as how to make use of these peptides as biopesticides or use knowl-
edge of their structures to invent new small-molecule toxophores that can
interact at the same target sites used by the peptides from the spider venom.
Chapters in this volume were chosen to provide a single comprehensive
review of structure and function of the insect midgut and the insecticidal
proteins and genes that have been used as alternatives to chemical insecti-
cides for controlling insect pests of agricultural and medical importance.
xii Preface
TARLOCHAN S. DHADIALLA
SARJEET S. GILL
CHAPTER ONE
Department of Molecular Microbiology and Immunology, The Johns Hopkins Bloomberg School of Public
Health, Baltimore, Maryland, USA
Contents
1. Introduction 1
2. Mosquito Larval Alimentary Canal 3
3. Other Insects 27
3.1 Lepidopteran larvae (caterpillars) 28
3.2 Coleopterans (beetles and their larvae) 30
3.3 Hemipterans (aphids) 31
4. Conclusions and Comment 32
References 33
Abstract
Insects as vectors of disease to humans and domesticated animals and as direct agri-
cultural pests are a source of tremendous economic and health-related challenge.
The eating habits of insects can provide the bases for disease transmission or the out-
right destruction of crops. The alimentary canal of insects is a common target and often
barrier for pest control strategies. Recent advances in technology have made it possible
to develop ever better understanding of the structure/function of the insect gut and
hence provide new and better targets for developing novel methods for limiting the
burdens that insects can present to humanity. In this review, we focus attention on
recent developments in our understanding of insect gut structure/function with partic-
ular emphasis on a few of the most challenging groups of insects: mosquitoes (dip-
terans), caterpillars (lepidopterans), beetles (coleopterans) and aphids (hemipterans).
1. INTRODUCTION
The alimentary canal of any higher organism is part of that organisms
first order environmental contact. Consequently insects have evolved highly
and natural competitors for the crops in the field. Relevant to this review of
course are a range of insect pests which consume or damage crops in a
variety of ways including the transmission of plant diseases. The impact of
pest insects on the world economy and the security of the human food sup-
ply is gigantic. Hence, we will also review the structural biology of certain
groups of agricultural threats: Lepidopteran and coleopteran larvae and
hemipteran life stages that impact crops. Of course, there are many and
diverse insects that will not be covered in this review but we hope to present
structural considerations that can be instructive and of fairly generalised
relevance.
Figure 1.1 Embryonic development of the main components of the alimentary canal in
insects. Invaginations of the ectoderm at the anterior and posterior poles give rise to the
foregut and hindgut, respectively. The midgut forms from endodermal cells adjacent to the
invaginations, proliferating and migrating to enclose the central yolk. The ectodermal and
endodermal tubes eventually fuse to form a contiguous alimentary canal. The sequence of
developmental steps progress from A through D. Redrawn by Gabriela Marie Ferguson
after several sources including Johansen and Butt (1941).
Figure 1.2 Architecture of the larval mosquito alimentary canal. Panel A: a scanning
electronmicrograph of a fourth instar larva of Aedes aegypti that has been dissected
to reveal the full length of the midgut amidst the exoskeleton and integument. Panel
B: a cross section of the gut epithelium in the region of the posterior midgut (PMG)
showing that it is a single cell thick epithelial tube but with varying morphological
characteristics of the individual cells. Panel C: a diagrammatic rendering of the larval
mosquito alimentary canal with labels applied for orientation. Numbers 18 indicate
abdominal segments. GC, gastric caecum; AMG, anterior midgut; CMG, central midgut;
PMG, posterior midgut; MT, malpighian tubules; HG, hindgut; Py, pyloris; Ai, anterior
intestine; Rc, rectum; Ac, anal canal; Ph, pharynx; Oe, oesophagus; Ca, cardia; pm,
peritrophic membrane; cm, caecal membrane; lm, gut lumen. The approximate pH of
the lumen of the gut is shown at the bottom of the cartoon. Taken from Linser et al.
(2007) with permission.
(SEM) of a fourth instar Aedes aegypti larvae (Linser et al., 2007). In this
image, the cuticular exoskeleton has been peeled back revealing the gross
architecture of nearly the entire alimentary canal. Figure 1.2B shows a
cross-sectional histological view that highlights one of the important char-
acteristics of the insect gut: it is a tubular organ system made up of a single cell
thick epithelium of highly polarised cells in terms of structure (i.e. apical,
lateral and basal structural distinctions) which presumably reflect distinctions
in function as well. Figure 1.2C provides a cartoon rendering of the major
organisational features or functional zones of the gut. A major subdivision of
the alimentary canal not shown in this cartoon is the pair of salivary glands
(SGs) that extend laterally from the oesophagus but these will enter the
6 Paul J. Linser and Rhoel R. Dinglasan
discussion later. Also, the structural components of the mouth and oral cav-
ity leading to the pharynx are not covered herein but the reader can find
many details in the literature (e.g. Clements, 1992). Figure 1.2C also depicts
one of the key structure/function relationships in many larval insect
alimentary canals: the lumen of the gut exhibits a range of pH values that
can reach extreme levels of basicity (Boudko et al., 2001; Corena et al.,
2004; Dadd, 1975; Terra et al., 1996; Zhuang et al., 1999). In mosquito lar-
vae, the anterior midgut (AMG) lumen pH can be as high as 11 (Boudko
et al., 2001; Corena et al., 2004; Dadd, 1975; Terra et al., 1996; Zhuang
et al., 1999). In certain other insect larvae such as caterpillars (e.g. Manduca
sexta) the luminal pH may exceed 12 (Cioffi, 1979; Harvey et al., 1983;
Wieczorek, 1992). The evolution of a digestive strategy that employs
extremely high pH is a subject of considerable interest both from the detailed
physiology of the system to the impact such pH extremes have on the imple-
mentation of control strategies that target gut function such as the bacterial
toxins from Bacillus thuringiensis and B. thuringensis israeliensis (BT and BTI
respectively; Gill et al., 1992; Chapter 2). This will be a recurring theme
within this review and volume.
The gross architecture of the larval mosquito gut is depicted in greater
detail in Fig. 1.3. We will address certain details for most of the specialised
functional zones. The first zone in this figure is the pair of bi-lobed (anterior
and posterior lobe) SGs. Although relatively little research has been done on
larval SGs, much is known about adult SGs as they are part of the infection
pathway for the transmission of viral and protozoan pathogens (Black and
Kondratieff, 2005). The SGs of larvae do in fact produce some of the earliest
Figure 1.3 Figure shows a detailed cartoon of key structural elements of the larval mos-
quito alimentary canal from the foregut including the salivary glands (SGs) to the rectum
(RC). Abbreviations are as in Fig. 1.2 except that the CMG is called the TR (transition
region) in this figure. Note the extent and location of the ectoperitrophic space
(ECTO), the cuticular lining of the foregut and hindgut and the variable distribution
and size of brush border membranes (microvilli) on the apical aspects of the gut cells
at the various regions of functional specialisation.
Insect Gut Structure, Function, Development and Target of Biological Toxins 7
(Seron et al., 2004). The basal aspect of both aforementioned types of caecal
cell exhibit varied but extensive plasma membrane infoldings (labyrinth),
again indicative of extensive expansion of the cell surface area. Figure 1.4
shows a representative transmission electron microscopic (TEM) image of
caecal cells from Aedes aegypti larva with these specific characteristics evident.
In addition to the two main types of caecal cell described above and numer-
ous times in the classical literature, a third type of epithlelial cell has been
noted. These cells typically occur at the posterior extreme of each caecal div-
erticulus and thus have been referred to as CAP cells (Seron et al., 2004).
CAP cells show very small or no microvilli on the apical surface and gen-
erally appear to be less broad in the apical to basal dimension. In subsequent
Figure 1.4 Gastric caeca cells as seen with transmission electron microscopy. Two dif-
ferent cells are shown: a lightly staining ion transporting cell (on the right) has micro-
villi (shown in cross section) containing mitochondria. The darker staining of the
resorptive cell (left) is due to the presence of extensive rough endoplasmic reticulum.
The scale bar represents 5 m. The inset is a high-magnification electron micrograph of
a transverse section from an ion-transporting cell with microvilli that contain mitochon-
dria. Portasomes (arrowheads) are prominent on the cytoplasmic face of the membrane.
Scale bar represents 100 nm. From Zhuang et al. (1999) with permission.
10 Paul J. Linser and Rhoel R. Dinglasan
the tube as shown here, which represents the AMG, has a very smooth
appearance. The lower portion, the PMG, shows a more granular surface.
At higher magnification or by viewing cross-sectioned material from the
AMG and PMG it is evident that the difference in appearance of the everted
tube is that the AMG cells have few and very short microvilli, whereas the
major cells of the PMG have apical arrays of microvilli that are tightly packed
and appear to be a solid cap when viewed by low magnification SEM (Clark
et al., 2005). At the region where the AMG and PMG meet, there is a tran-
sitional region that is several cells broad from anterior to posterior along the
12 Paul J. Linser and Rhoel R. Dinglasan
long axis of the gut. The apical surfaces of cells in the transitional region
show an increase in the numbers and dimensions of microvilli (Clark
et al., 2005). Other distinctions have been described between the AMG,
transitional region and the PMG and we will return to these shortly.
The PMG is broader than the AMG and the epithelial cells are also larger
than those of the AMG. The apical surface is tremendously extended by
microvilli that are tightly packed. Ultrastructural analyses show basal mem-
brane infoldings, copious numbers of intracellular vesicles and extensive
mitochondrial profiles, which are indicative of cells that are very metabol-
ically active and involved in absorptive and secretory processes (Billingsley,
1990; Clements, 1992; Zhuang et al., 1999).
The hindgut is composed of the pyloris and MTs, ileum (anterior intes-
tine), rectum and anal canal that are all of ectodermal origin in distinction
with the endodermal midgut. The PM, which originates at the cardia at
the posterior end of the foregut, begins to lose its integrity as it enters the
pyloris of the hindgut (HG). The HG has a cuticular lining that encompasses
the PM and the food bolus as it continues its journey toward excretion. The
cells of the pyloris are thin epithelial cells and the funnel-shaped region of the
HG has a posterior band of muscle forming a sphincter (the pyloric sphinc-
ter). At the most anterior extreme of the pyloris, the epithelial cells are quite
small and contiguous with the stem cells of the posterior imaginal ring
(Clements, 1992; Klowden, 2007). The five MTs are tubes with an opening
into the pyloris and a closed terminus at the distal end of each. Two types of
cells are described in the MTs: principal cells and stellate cells, which have
differential embryological origins (Davies and Terhzaz, 2009; Dow, 2009).
Principal cells are large with extensive apical microvilli and each cell can
extend nearly around the circumference of the MT lumen, which is zigzag
shaped due to the apical intrusion of the principal cells. Principal cells possess
a large polytene nucleus and in later stages of larval development accumu-
lations of membrane-bound inclusions or concretion bodies (Bradley and
Snyder, 1989). The second type of cell in the MT is called the stellate cell
and their shape is indeed star like. Stellate cells localise between some of the
principal cells and can be difficult to detect with simple microscopy as they
can appear to be the interconnections of the lateral membranes of principal
cells. Recent physiological and immunohistochemical analyses have pro-
vided insights into the structural relationships between principal cell and
stellate cells (see later discussion of this section). Stellate cells possess much
smaller nuclei than principal cells and it is not clear whether or not they are
diploid or polyploid. The function of the MTs is generally held to be similar
Insect Gut Structure, Function, Development and Target of Biological Toxins 13
to that of the vertebrate kidney in ion regulation and the formation of the
primary urine (Beyenbach et al., 2009; Bradley, 1987).
The ileum or anterior intestine is a thin-walled epithelium within a
prominent muscular tube. Relatively little is known of the functional spe-
cialties of the ileum other than it serves as the continuation of the pathway
for the movement of the excreta. The muscular tube surrounding the ileum
is at least partly responsible for pushing faecal material through the rectum
and into the anal canal.
The rectum is a complex region of the gut and exhibits substantial archi-
tectural variations between genera of mosquitoes and insects in general. Sim-
ply stated it is an essential organ in the regulation of ionic balance in the
animal and the retention or elimination of specific solutes. The cells of
the insect rectum have at times been described as among the most complex
cells in biology (Berridge and Oschman, 1972). The rectum of mosquito
larvae varies in general architecture. The tubular nature of the alimentary
canal continues into the ectodermal rectum. The length and architecture
of the rectum varies between species but, generally, the rectum is a
cuticle-lined epithelium that can exhibit longitudinal folds. The two sub-
families of mosquitoes, the Culicinae and the Anophelinae are somewhat
different in terms of rectum structure (Bradley, 1987; Smith et al., 2007,
2008, 2010; White et al., 2013). Additionally, the osmolarity of the aquatic
environment in which larvae develop can be reflected in sometimes-gross
variations on the architectural details of the rectum. Culicinae species, such
as Aedes aegypti, which select low ionic strength aquatic habitats exhibit a
single compartment rectum. In contrast, Culicinae, such as Aedes
campestris, that are tolerant of much higher osmolarities possess a rectum with
distinct anterior and posterior compartments (Bradley, 1987; Smith et al.,
2007, 2008, 2010; White et al., 2013). Anophelinae such as Anopheles
gambiae (salt intolerant) and Anopheles merus (salt tolerant) possess rectal struc-
tures that are somewhat intermediate between the partitioned (anterior and
posterior) rectum of salt-tolerant culicinae and the single-compartment rec-
tum of salt intolerant species. Specifically, Anophelinae possess two distinct
populations of cells but lack a clear compartmental divide between them
(Bradley, 1987; Smith et al., 2007, 2008, 2010; White et al., 2013). The
two populations are functionally and structurally distinct and have been
named the dorsal anterior rectum (or DAR) cells and the remaining non-
DAR cells (Smith et al., 2007). We will return to a discussion of the struc-
ture/function details of these cells later. In general, whether discussing Cul-
icinae or Anophelinae larval rectum cells, the cells are characterised by
14 Paul J. Linser and Rhoel R. Dinglasan
ectoperitrophic space, which is in contact with the lumen of the gut and the
lumen of the caeca. Various subdivisions of the gut are associated with char-
acteristic structural specialisation of the several cell types located therein. In
relation to the accessibility and effectiveness of orally administered biological
toxins (a focus of this volume), gut structure is of obvious importance. In
addition, certain details of the cell biology that have come to light in recent
years have either already become demonstrably important to new control
strategies or clearly have that potential. At this point, we are going to review
a number of fairly recent investigations that have provided insight into the
structure/function of the important cell types of the alimentary canal.
Material ingested by mosquito larvae will encounter the secretions of the
SGs (i.e. saliva) as one of the earliest steps in digestion. In addition to aiding
digestion, the saliva contains numerous gene products associated with
immune surveillance and the inactivation of potential pathogens. Trans-
criptomic and proteomic analyses of the Anopheles gambiae larval SGs rev-
ealed the production and secretion of such immune effectors as defensins,
lysozyme and TIL-domain proteins (Neira Oviedo et al., 2009). Hence,
ingestion of materials that might inactivate such components of the saliva
may well render the larvae more susceptible to biological toxins or organ-
isms. Methods for generally inhibiting SG function would also be reasonable
targets for the development of novel control strategies. The anterior and
posterior lobes of the SGs are biochemically distinguishable but nothing
is known about the actual compartmentalisation of specific salivary compo-
nent synthesis and secretion. A clearer understanding of SG structure/func-
tion and cell biology will be a valuable pursuit in the future.
The glycocalyx-type extracellular matrix linings of the alimentary canal
provide a range of functions such as facilitating the one-way movement of
the food bolus from the mouth to the anus, physical protection of
delicate cell surfaces from what can be abrasive particulates, barriers to
full blown biological invasion from microbiota in the food bolus, size-
exclusion barriers to macromolecular diffusion and even selective perme-
ability and toxin sequestration. The lining of the foregut and the hindgut
is cuticular and exhibits structural qualities of cuticle. The PM and the
CM are also chitin-containing acellular barrier matrices with diverse func-
tions (Hegedus et al., 2009; Lehane, 1997; Rudin and Hecker, 1989). The
CM and PM are distinguishable from each other as well as the cuticular lin-
ings of the foregut and hindgut on several structural and biochemical bases.
For example, lectin labelling shows that the CM and PM are readily labelled
with several lectins including wheat germ agglutinin but the cuticular
16 Paul J. Linser and Rhoel R. Dinglasan
exoskeleton and the linings of the foregut and hindgut are distinguished
from the PM and CM with ricinus communis 1 (RCA; Linser et al.,
2008; Neira Oviedo et al., 2009). Figure 1.6 shows such a comparison
between Dolichus biflorus Agglutinin (DBA) and RCA1 in a longitudinal sec-
tion of a fourth instar Anopheles gambiae larva. The green fluorescence (DBA)
highlights the PM and CM vividly and in contrast to the red signal (RCA1),
which stains the exoskeleton cuticle and the linings of the foregut and hind-
gut. Indeed one can see that in the pyloris, the red cuticular lining of the
hindgut surrounds the PM and that the PM begins to be compacted at
the junction with the ileum. Furthermore, the origin of the PM (in green)
is visible in the folded layering of the cardia while the foregut cuticle is visible
within the innermost channel of the termination of the oesophagus (as col-
ours are not presented in the print version of this volume, please visit the
on-line version for full colour details). These chitin-containing extra cellular
matrices (ECMs) are structural barriers to any ingested material and hence
need to be penetrated by any biological toxins or control organisms. Lectins,
as used herein, are capable of discriminating specific details of glyc-
oconjugate structure in terms of specific sugar structure, chemical linkages
and other details. The analyses described here serve to show that the detailed
glycobiology of the ECMs associated with gut possess distinct biochemical
signatures. The impact of that varied biochemistry is largely untested.
The structure of specific cell types within the gut provides insight into
their functional roles. As described earlier, the apical surface of the large,
principal epithelial cell types of the GC and PMG possess extensive arrays
of microvilli. This implies an absorptive role. AgAPN1, a GPI-anchored
plasma membrane glycoprotein, first identified in adult Anopheles gambiae
is a peptidase presumably involved in the final stages of protein digestion,
liberating amino acids for absorption (Dinglasan et al., 2007). AgAPN1 is
also a putative point of attachment for the malaria parasite Plasmodium
falciparum (Armistead et al., 2014; Dinglasan et al., 2007; Mathias et al.,
2013). Antibodies to this protein specifically label the microvillar arrays
on the PMG cells and on a specific subset of the GC cells. Figure 1.7 shows
AgAPN1 labelling of the PMG and the GC cells that form the neck region of
each GC lobe. Only GC cells that are exterior to the CM label for this pro-
tein. In contrast, AgAPN2, another distinct cell surface aminopeptidase
thought to be a binding site for the Cry11Ba toxin of Bacillus thuringiensis
(Zhang et al., 2008) is also found on the microvilli of PMG cells and GC
cells. But, in this case, only the GC cells that lie within the GC, internal
to the CM, exhibit the protein. Additionally, the CAP cells (see earlier dis-
cussion) at the posterior extreme of each caecum contrast with the surround-
ing neighbouring GC cells by lacking AgAPN2 (Fig. 1.8; Harvey et al.,
2010; Linser et al., 2007).
As mentioned earlier, one of the striking qualities of the larval mosquito
alimentary canal is the extreme pH of portions of the gut lumen (Fig. 1.2).
The evolutionary selective pressure for extremely high pH in the gut lumen
is often associated with the high content of plant material in the diets of many
insect larvae including caterpillars and mosquito larvae (Terra et al., 1996).
This may have some truth but it should be noted that even in the Tsetse
(Glossinidae) which derives all of its biological energy for its entire life cycle
from blood meals, the gut pH can exceed 10 (Liniger et al., 2003). Regard-
less of the physiological ramifications of an alkaline digestive system, the
mechanisms which drive the pH gradient along the length of the mosquito
larva gut from nearly neutral at the level of the GC, to pH 10.5 or even
18 Paul J. Linser and Rhoel R. Dinglasan
Figure 1.7 Figure shows the distribution of two integral membrane amino peptidases,
AgAPN1 (panel A) and AgAPN2 (panel B) in larval Anopheles gambiae gut sections. In (A),
AgAPN1 (blue) is seen on the brush border membranes (BBM) of the posterior midgut
(PMG) cells (short arrow to left) and the BBM of the neck of the gastric caeca (GC) (short
arrow to right and in high mag images at bottom of panel). Note that FITC-conjugated
Vicia Villosa Lectin (green) was used to highlight the cuticular structures, the peritrophic
membrane (PM) and the caecal membrane (CM) (long arrows) making the limitation of
AgAPN1 to the neck cells of the GC evident. In (B), AgAPN2 (green) is compared to Na+/
K+-ATPase (red) and the cytoplasmic marker carbonic anhydrase-9 (CA9). The short
arrows indicate labelling for AgAPN2 on the BBM of the PMG and on the BBM of the
GC cells internal to the CM. From Linser et al. (2008) with permission.
Insect Gut Structure, Function, Development and Target of Biological Toxins 19
Figure 1.8 Figure shows the GC of fourth instar Anopheles gambiae at high magnifica-
tions to highlight the distribution of AgAPN2 (green) relative to the basal membrane
marker Na+/K+-ATPase (red) and the cytoplasmic CA9 (blue). Panels A and B show
the anterior portion of a caecum at the junction with the AMG. Note that AgAPN2 is
prominently localised to the BBM of the anterior GC cells. Panels C and D show the
posterior portion on the same caecum. Note the complete loss of staining for both
Na+/K+-ATPase and AgAPN2 (arrow) at the posterior extreme of the caecum, which is
the CAP cells. The inset in D shows only CA9 staining at the same magnification as
shown in A and B to provide reference. From Linser et al. (2008) and Harvey et al.
(2010) with permission.
higher in the AMG, to pH 8 in the PMG and slightly acidic pH in the rec-
tum are a study in functional cell polarity (Filippova et al., 1998; Zhuang
et al., 1999). To clarify, the major process that drives the up and down
pH gradient is the pumping of protons via a proton pump called
vacuolar-ATPase (V-ATPase; Filippova et al., 1998; Zhuang et al., 1999).
V-ATPase and its roles in epithelial physiology have been studied exhaus-
tively in numerous model systems. For the purpose of our structural discus-
sion here, pharmacological studies have demonstrated a central role in
establishing and maintaining the highly alkaline environment within the
gut. The V-ATPase is actually a macromolecular complex of several proteins
20 Paul J. Linser and Rhoel R. Dinglasan
Figure 1.9 Figure demonstrates the apical to basal back to apical transitions in the
polarised distribution of V-ATPase in the Aedes aegypti larval midgut. Panels AC show
regions of the GC, AMG and PMG, respectively. Green/yellow labelling shows the
localisation of V-ATPase on the apical BBM of the GC (A) and PMG (C) and on the basal
membrane of the AMG (B). Red labelling identifies the basal side of the epithelium in all
cases as it depicts the external layer of muscle labelled with TRITC-conjugated Phalloidin.
Blue is DAPI staining of cell nuclei. Panels DF show the same series of gut regions at
higher magnification. From Zhuang et al. (1999) with permission.
the transitional region (TR; Clark et al., 2005; Smith et al., 2007). The cells
of the TR exhibit a graded shift in several parameters including the numbers
and size of microvilli, susceptibility to Cry4Ba toxin damage, the patterning
of cell nuclei, and the flip-flop polarised location of the V-ATPase and NaK-
ATPase (Clark et al., 2005; Smith et al., 2007). Figure 1.10 shows an analysis
of this region of the Anopheles gambiae larval midgut with immunohisto-
chemical markers for NaK-ATPase. Panel C provides a view of the shift
in functional polarity with NaK-ATPase on the basal membrane infoldings
in the PMG (to the right) and the shift of this protein to the apical side in the
AMG cells (to the left; Smith et al., 2007). Also shown in this figure is the
distribution of another important enzyme in the regulation of gut and cel-
lular pH, carbonic anhydrase 9 (CA9) which is 1 of 12 mosquito carbonic
anhydrase genes/proteins. This pH regulator is expressed by cells of the GC
and the rectum as a cytoplasmic protein but is also secreted by the GC cells
22 Paul J. Linser and Rhoel R. Dinglasan
Figure 1.10 Figure shows the entire length of the larval Anopheles gambiae alimentary
canal from the cardia at the extreme to the rectum at far right in panel A. CA9 is labelled
in green and Na+/K+-ATPase in red. The arrow in (A) indicates the transition region at
which Na+/K+-ATPase shifts from basal in the AMG to apical in the PMG. Panels BD
show the transition region at higher magnification. CA9 is expressed by several cell
types but is secreted by GC cells into the ectoperitrophic space. In the transition region,
CA9 is also found in cell nuclei (arrows, B and D). Panel C isolates the Na+/K+-ATPase
signal and thus reveals the shift in epithelial cell polarity from basal (PMG, solid arrow)
to apical (AMG, hollow arrow) of this very important physiological function. Panel
D overlays the red and green images with a blue (DRAQ5, DNA) signal. From Smith
et al. (2007) with permission.
into the ectoperitrophic space and hence part of the digestive milieu and a
component that will be encountered by ingested microbiota and toxins.
The PMG of larval mosquitoes is the largest region of the alimentary
canal in terms of cell surface exposure to ingested materials. The apical brush
border of the principal PMG cells is massive and difficult to estimate in terms
of gross quantity. Its extensive nature has made it possible to engineer rather
Insect Gut Structure, Function, Development and Target of Biological Toxins 23
Figure 1.11 Figure reveals several details of MT cell biology in the hindgut region of
Anopheles gambiae larva sections. Panel A shows immunolabelling for the bicarbonate
transporter AgNDAE1 (red), V-ATPase (green) and CA9 (blue). The MT is characterised by
basal membrane staining of the principal cells for AgNDAE1 and internal and apical
staining for V-ATPase. Panel A also shows a partial view of the rectum and that CA9
is characteristic of the DAR cells whereas V-ATPase is also present in and on the non-
DAR cells. Panels B and C show a high magnification view of a cross section through
an MT displaying the basal labelling for AgNDAE1 and the apical labelling for
V-ATPase. The arrow in (A) indicates a stellate cell which shows little signal in this label-
ling combination. Panels DH show high magnification views of MT whole mounts
labelled for proton antiporter NHA2 (red), Na+/K+-ATPase (blue) and Griffonia
simplicifolia I lectin (green) and in D&H, DAPI for nuclei (aqua). NHA2, Na+/K+-ATPase
and GSL-I all label the stellate cells of the MT. In the merge images G and H, the polarity
of the markers is revealed. NHA2 and GSL1 co-localise (long arrow in H) to the apical
aspect of the stellate cell whereas Na+/K+-ATPase is basal. Figures AC from Linser
et al. (2012) and DH from Xiang et al. (2012) with permission.
Insect Gut Structure, Function, Development and Target of Biological Toxins 25
number of cells) and much smaller than principal cells. Stellate cells are first
evident amongst the principal cells after the first (proximal to the hindgut)
10% of the length of each MT. Like the principal cells, stellate cells exhibit
specific patterns of plasma lemmal proteins on both the apical and basal/
lateral membranes (Beyenbach, 2003; Linser et al., 2012; Xiang et al.,
2012). Models for the physiological roles of both principal cells and stellate
cells have been proposed and supported by many varied studies (Beyenbach,
2003; Linser et al., 2012; Xiang et al., 2012). An interesting and controver-
sial characteristic of MTs is that a substantial series of investigations supports
the movement of certain ions such as Cl between cells, which indicates a
certain specialisation of cellcell junctions that can provide a regulated peri-
cellular pathway (Beyenbach, 2003). In general, MTs monitor the makeup
of the hemolymph and actively produce the primary urine, providing one of
the key homeostatic functions within the larva.
The final component of the larval alimentary canal to be discussed herein
is the rectum. As mentioned earlier, the rectum is contiguous with the ante-
rior components of the alimentary canal by the ileum or posterior intestine.
The rectum is a major regulator of excretion, elimination and retention of
solutes critical for homeostasis. Mosquito larvae have adapted to many
aquatic environments that can range in ionic strength from very low salinity
fresh water to salt concentrations that exceed that of sea water (Smith et al.,
2010; White et al., 2013). One rather obvious structural variation that cor-
relates with salt tolerance is the structure of the rectum. Figure 1.12 shows
the three major generalised structures of rectum from the simple bulb made
up of seemingly a single type of cell found in culicine mosquitoes with low
salt tolerance, to the bi-lobed rectum of salt-tolerant culicines to the region-
ally specialised two-cell type form in anopheline mosquito larvae (Smith
et al., 2007, 2008). The salt tolerance of mosquito larvae has been investi-
gated extensively and species that are capable of short term adaptation and
those that have very narrow ranges of salinity are well documented in the
classical literature (e.g. Bradley, 1987; Clements, 1992). Recent investiga-
tions have also shown that certain species of anopheline mosquitoes exhibit
dynamic structural changes in the rectum when placed into changing con-
ditions of salinity (Smith et al., 2008; White et al., 2013). It appears that the
balance of functionalities compartmentalised within the two rectum cell
types, DAR and non-DAR can change in response to ionic fluctuations.
An interpretation of this dynamic is that the associations between transport
mechanisms compartmentalised within the cell types can be modified to
produce different potentials for vectorial and linked ion transport (Smith
26 Paul J. Linser and Rhoel R. Dinglasan
Figure 1.12 Figure depicts the gross structure of the three types of larval mosquito rec-
tum as seen in fresh-water culicenes, salt-tolerant culicenes and all anophelines. Fresh-
water culicenes such as Aedes aegypti have a uniform rectum with one primary cell type
as seen in this whole mount immunolabelled for basal NaK-ATPase (left). Salt-tolerant
culicenes such as Ochlerotatus taeniorhynchus have a rectum with distinct anterior and
posterior regions. Anophelines possess a rectum that is distinguished by an anterior
patch of distinct cells termed dorsal anterior rectum (DAR) cells and a posterior majority
of non-DAR cells. AR, anterior rectum; PR, posterior rectum. Image modified from Smith
et al. (2008).
Figure 1.13 Figure shows whole mount preparations of isolated midgut from fourth
instar larvae of Aedes aegypti (AC) and Anopheles gambiae (DF) immunolabelled for
the peripheral membrane protein carbonic anhydrase CA-10 (green) and Phalloidin to
reveal muscle (red). The arrows in (A) and the merged images show that a specific subset
of the gut musculature labels for CA-10 whereas other muscles of the gut do not. Thus the
musculature of the gut exhibits more structure/function variability than simple longitudi-
nal versus circumferential. From Seron et al. (2004) with permission.
3. OTHER INSECTS
The general architecture of the alimentary canal of insects varies con-
siderably depending on adaptive specialisations. A useful summary is
28 Paul J. Linser and Rhoel R. Dinglasan
Figure 1.14 Structure of the lepidopteran midgut. Regional specialisation into three
zones is visible at both coarse and progressively finer resolution. Approximate magni-
fications are shown. At 500 , the two forms of goblet cell are evident in the regions of
the gut epithelium. Taken from Dow (1992) with permission.
among the columnar cells are the very distinct goblet cells. These cells are so
named because of their architecture, which resembles a goblet with a prom-
inent internal cavity. Goblet cells of the AMG are somewhat different from
those of the PMG in that in the AMG goblet cells the goblet cavity is prox-
imal to the basal aspect of the cell whereas in the PMG goblet cells the cavity
is adjacent to the apical extreme of the cell (Fig. 1.14). The inner surface of
the goblet cell plasma membrane adjacent to the cavity is studded with
potasomes as described earlier in the mosquito system. The lepidopteran lar-
val gut, due to its large size and its tractability as an experimental system,
made it possible for numerous investigators to track down the nature of
the portasome and to demonstrate the role that the V-ATPase plays in insect
epithelial physiology (Harvey et al., 1983; Klein et al., 1991) and as a target
for novel control strategies (Baum Chapter 5 this volume). The activity of
this cell-surface proton pump is the driving force of another striking char-
acteristic of the lepidopteran gut: a luminal pH that can be as high as 12 and
transepithelial potential of nearly 240 mV (Harvey et al., 1983; Wieczorek,
30 Paul J. Linser and Rhoel R. Dinglasan
1992). These are physiological extremes found nowhere else in nature. Cer-
tainly part of capacity for lepidopterans to generate these remarkable phys-
iological conditions is contributed by the unique structure/function of the
epithelium and in particular the goblet cells (Harvey et al., 1983;
Wieczorek, 1992).
Lepidopterans do possess MTs although in the vicinity of the rectum
they differ from mosquitoes. The distal ends of the MTs in lepidopterans
(and coleopterans) are embedded in an extracellular matrix that holds them
in a fixed juxtapositional relationship with the basal surface of the rectum
(Azuma et al., 2012; Fermino et al., 2010; Ramsay, 1976). This association
between the tissues is called the cryptonephric rectal complex (Azuma et al.,
2012; Fermino et al., 2010; Ramsay, 1976). Most other insects have free dis-
tal ends that can actually be moved about the hemocoel by muscular con-
tractions in the body wall. The physiological details of this fixed association
with the rectum are poorly understood but is the subject of numerous inves-
tigations (Azuma et al., 2012; Fermino et al., 2010; Ramsay, 1976).
REFERENCES
Al-Deeb, M., Wilde, G., 2005. Effect of Bt corn expressing Cry3Bb1 toxin on western corn
rootworm (Coleoptera: Chrysomelidae) biology. J. Kansas Entomol. Soc. 78, 142152.
Armistead, J.S., Morias, I., Mathias, D.K., Jardim, J.G., Joy, J., Fridman, A., Finnefrock, A.C.,
Baguchi, A., Plebanski, M., Scorpio, D.G., Churcher, T.S., Borg, N.A., Sattabongkot, J.,
Dinglasan, R.R., et al., 2014. Antibodies to a single, conserved epitope in Ahopheles
APN1 inhibit universal transmission of Plasmodium falciparum and Plasmodium vivax
malaria. Infect. Immun. 82, 818829.
Azuma, M., Nagae, T., Maruyama, M., Kataoka, N., Miyake, S., 2012. Two water-specific
aquaporins at the apical and basal plasma membranes of insect epithelia: molecular basis
for water recycling through the cryptonephric complex of lepidopteran larvae. J. Insect
Physiol. 58, 523533.
Berridge, M.J., Oschman, J.L., 1972. Insect Epithelia. In: Berridge, M.J., Oschman, J.L.
(Eds.), Transporting Epithelia. Academic Press, New York, NY, pp. 5860.
Beyenbach, K.W., 2003. Transport mechanisms of diuresis in Malpighian tubules of insects.
J. Exp. Biol. 206, 38453852.
Beyenbach, K.W., Baumgart, S., Lau, K., Piermarini, P.M., Zhang, S., 2009. Signalling to
the apical membrane and to the paracellular pathway: changes in the cytosolic proteome
of Aedes Malpighin tubules. J. Exp. Biol. 212, 329340.
Billingsley, P.F., 1990. The midgut ultrastructure of hematophagus insects. Annu. Rev.
Entomol. 35, 219248.
Billingsley, P.F., Lehane, M.J., 1996. Structure and ultrastructure of the insect midgut.
In: Lehane, M.J., Billingsley, P.F. (Eds.), Biology of the Insect Midgut. Chapman &
Hall, London, pp. 330.
Black, W.C., Kondratieff, B.C., 2005. Evolution of arthropod disease vectors.
In: Marquardt, W.C. (Ed.), Biology of Disease Vectors, second ed. Elsevier Academic
Press, Burlington, MA, pp. 923.
Boudko, D.Y., Harvey, W.R., Linser, P.J., 2001. Alkalinization by chloride/bicarbonate
pathway in larval mosquito midgut. Proc. Natl. Acad. Sci. U. S. A. 98, 1535415359.
Bradley, T.J., 1987. Physiology of osmoregulation in mosquitoes. Annu. Rev. Entomol.
32, 439462.
Bradley, T.J., Snyder, C., 1989. Fluid secretion and microvillar ultrastructure in mosquito
Malpighian tubules. Am. J. Physiol. 257, R1096R1102.
Bragard, C., Caciagli, P., Lemaire, O., Lopez-Moya, J.J., MacFarlane, S., Peters, D., Susi, P.,
Torrance, L., 2013. Status and prospects of plant virus control through interference with
vector transmission. Annu. Rev. Phytopathol. 51, 177201.
34 Paul J. Linser and Rhoel R. Dinglasan
Brault, V., Uzest, M., Monsion, B., Jacquot, E., Blanc, S., 2010. Aphids as transport devices
for plant viruses. C. R. Biol. 333, 524538.
Brown, M.R., Lea, A.O., 1988. FMRFamide- and adipokinetic hormone-like immunore-
activity in the nervous system of the mosquito, Aedes aegypti. J. Comp. Neurol.
270, 606614.
Chagas, A.C., Calvo, E., Rios-Velasquez, C.M., Pessoa, F.A., Madeiros, J.F., Ribeiro, J.M.,
2013. A deep insight into the sialotranscriptome of the mosquito Psorophora albipes. BMC
Genomics 14, 875894.
Chougule, N.P., Bonning, B.C., 2012. Toxins for transgenic resistance to hemipteran pests.
Toxins 4, 405429.
Chougule, N.P., Li, H., Liu, S., Linz, L.B., Narva, K.E., Meade, T., Bonning, B.C., 2012.
Retargeting of the Bacillus thuringiensis toxin Cyt2Aa against hemipteran insect pests.
PNAS 110, 84658470.
Chu, C., Spencer, J.L., Curzi, M.J., Zavala, J.A., Seufferheld, M.J., 2013. Gut bacteria facil-
itate adaptation to crop rotation in the western corn rootworm. PNAS
110, 1191711922.
Cioffi, M., 1979. The morphology and fine structure of the larval midgut of a moth (Manduca
sexta) in relation to active ion transport. Tissue Cell 11, 467479.
Clark, T.M., Hutchinson, M.J., Huegel, K.L., Moffett, S.B., Moffett, D.F., 2005. Additional
morphological and physiological heterogeneity within the midgut of larval Aedes aegypti
(Diptera: Culicidae) revealed by histology, electrophysiology, and effects of Bacillus thur-
ingiensis endotoxin. Tissue Cell 37, 457468.
Clements, A.N., 1992. The Biology of Mosquitoes. vol. 1. Chapman & Hall, London,
pp. 100149.
Corena, M.P., Fiedler, M.M., VanEkeris, L., Linser, P.J., 2004. A comparative study of carbonic
anhydrase in the midgut of mosquito larvae. Comp. Biochem. Physiol. 137, 207225.
Dadd, R.H., 1975. Alkalinity within the midgut of mosquito larvae with alkaline-active
digestive enzymes. J. Insect Physiol. 21, 18471853.
Davies, S.A., Terhzaz, S., 2009. Organellar calcium signaling mechanisms in Drosophila epi-
thelial function. J. Exp. Biol. 212, 387400.
Dedryver, C., Le Ralec, A., Fabre, F., 2010. The conflicting relationships between aphids
and men: a review of aphid damage and control strategies. C. R. Biol. 333, 539553.
Dinglasan, R.R., Kalume, D.E., Kanzok, S.M., Ghosh, A.K., Muratova, O., Pandey, A.,
Jacobs-Lorena, M., 2007. Disruption of Plasmodium falciparum development by anti-
bodies against a conserved mosquito midgut antigen. PNAS 104, 1346113466.
Dow, J.T., 1992. pH gradients in lepidopteran midgut. J. Exp. Biol. 172, 355375.
Dow, J.T., 2009. Insights into the Malpighian tubule from functional genomics. J. Exp. Biol.
212, 435445.
Edwards, M.J., Jacobs-Lorena, M., 2000. Permeability and disruption of the peritrophic
matrix and caecal membrane from Aedes aegypti and Anopheles gambiae mosquito larvae.
J. Insect Physiol. 46, 13131320.
Engel, P., Moran, N.A., 2013. The gut microbiota of insectsdiversity in structure and
function. FEMS Microbiol. Rev. 37, 699735.
EPA (Environmental Protection Agency), 2003. Event MON863 BtCry3Bb1 corn biopes-
ticide registration action document. Available free at: http://www/epa.gov/pesticides/
biopesticides/ingredients/tech_docs/brad_006484.htm.
Fermino, F., Conte, H., Falco, J.R.P., 2010. Analysis of nucleus activity in Malpighian
tubules of Diatrea saccharalis (Fabricus) (Lepidoptera: Crambidae) larvae by critical elec-
trolyte concentration. Neotrop. Entomol. 39, 568571.
Filippova, M., Ross, L.S., Gill, S.S., 1998. Cloning of the V-ATPase B subunit cDNA from
Culex quinquefasciatus and expression of the B and C subunits in mosquitoes. Insect Mol.
Biol. 7, 223232.
Insect Gut Structure, Function, Development and Target of Biological Toxins 35
Frank, D.L., Bukowsky, R., French, B.W., Hibbard, B.E., 2011. Effect of MIR604 trans
genic maize at different stages of development on western corn rootworm
(Coleoptera: Chrysomelidae) in a central Missouri field environment. J. Econ. Entomol.
104, 20542061.
Gill, S.S., Cowles, E.A., Pietrantonio, P.V., 1992. The mode of action of Bacillus thuringiensis
endotoxins. Annu. Rev. Entomol. 37, 615636.
Goltsev, Y., Rezende, G.L., Vranizan, K., Lanzaro, G., Valle, D., Levine, M., 2009. Devel-
opmental and evolutionary basis for drought tolerance of the Anopheles gambiae embryo.
Dev. Biol. 330, 462470.
Harvey, W.R., Cioffi, M., Dow, J.A.T., Wolfersberger, M.G., 1983. Potassium ion transport
ATPase in insect epithelia. J. Exp. Biol. 106, 91117.
Harvey, W.R., Boudko, D.Y., Rheault, M.R., Okech, B.A., 2009. NHEvnat: an H +
V-ATPase electrically coupled to a Na+:nutrient amino acid transporter (NAT) forms
a Na +/H+ exchanger (NHE). J. Exp. Biol. 212, 347357.
Harvey, W.R., Okech, B.A., Linser, P.J., Becnel, J.J., Ahearn, G.A., Sterling, K.M., 2010.
H+-V-ATPase-energized transporters in brush border membrane vesicles from whole
larvae of Aedes aegypti. J. Insect Physiol. 56, 13771389.
Hegedus, D., Erlandson, M., Gillott, C., Toprak, U., 2009. New insights into peritrophic
matrix synthesis, architecture, and function. Annu. Rev. Entomol. 54, 285302.
Hogenhout, S.A., Ammar, E.D., Whitfield, A.E., Redinbaugh, M.G., 2008. Insect
vector interactions with persistently transmitted viruses. Annu. Rev. Phytopathol.
46, 327359.
Jeffers, L.A., Roe, R.M., 2008. The movement of proteins across the insect and tick digestive
system. J. Insect Physiol. 45, 319332.
Johansen, O.A., Butt, F.H., 1941. Embryology of Insects and Myriopods. McGraw Hill,
New York, NY.
Klein, U., Loffelmann, G., Wieczorek, H., 1991. The midgut as a model system for insect
K+-transporting epithelia: immunocytochemical localization of a vacuolar-type H +
pump. J. Exp. Biol. 161, 6175.
Klowden, M.J., 2007. Physiological Systems in Insects, second ed. Academic Press,
Burlington, MA, pp. 137179.
Krajniak, K.G., 2005. Annelid endocrine disruptors and a survey of invertebrate
FMRFamide-related peptides. Integr. Comp. Biol. 45, 8896.
Lehane, M.J., 1997. Peritrophoci matrix structure and function. Annu. Rev. Entomol.
42, 525550.
Lehane, M.J., Billingsley, P.F., 1996. Biology of the Insect Midgut, first ed. Chapman & Hall,
London, pp. 325.
Liniger, M., Acosta-Serrano, A., Van Den Abbeele, J., Renggli, C.K., Brun, R., Englund, P.T.,
Roditi, I., 2003. Cleavage of trypanosome surface glycoproteins by alkaline trypsin-like
enzyme(s) in the midgut of Glossina morsitans. Int. J. Parasitol. 33, 13191328.
Linser, P.J., Boudko, D.Y., Corena, M.P., Harvey, W.R., Seron, T.J., 2007. The molecular
genetics of larval mosquito biology. J. Am. Mosq. Control Assoc. 23 (2 Suppl.), 283293.
Linser, P.J., Neira-Oviedo, M.V., Smith, K.E., 2008. Glycoconjugate analysis in Anopheles
gambiae. Am. J. Trop. Med. Hyg. 79 (Suppl. 6), 150199.
Linser, P.J., Neira Oviedo, M., Hirata, T., Seron, T.J., Smith, K.E., Piermarini, P.M.,
Romero, M.F., 2012. Slc4-like anion transporters of the larval mosquito alimentary
canal. J. Insect Physiol. 58, 551562.
Marinotti, O., Calvo, E., Nguyen, Q.K., Dissanayake, S., Ribiero, J.M., James, A.A., 2006.
Genome-wide analysis of gene expression in adult Anophleles gambiae. Insect Mol. Biol.
15, 112.
Mathias, D.K., et al., 2013. A small molecule glycosaminoglycan mimetic blocks Plasmodium
invasions of the mosquito midgut. PLoS Pathog. 9, e1003757.
36 Paul J. Linser and Rhoel R. Dinglasan
Matter, K., Balda, M.S., 2003. Signalling to and from tight junctions. Mol. Cell. Biol.
4, 225236.
Meredith, J., Phillips, J.E., 1973. Rectal ultrastructure in salt-and freshwater mosquito larvae
in relation to physiological state. Z. Zellforsch. Mikrosk. Anat. 138, 122.
Neira Oviedo, M., Ribeiro, J., Heyland, A., VanEkeris, L., Moroz, T., Linser, P.J., 2009.
The salivary transcriptome of Anopheles gambiae (Diptera: Culcidae) larvae: a
microarray-based analysis. Insect Biochem. Mol. Biol. 39, 382394.
Ng, J.C., Perry, K.L., 2004. Transmission of plant viruses by aphid vectors. Mol. Plant Pat-
hol. 5, 505511.
Patrick, M.L., Aimanova, K., Sanders, H.R., Gill, S.S., 2006. P-type Na+/K+-ATPase and
V-type H+-ATPase expression patterns in the osmoregulatory organs of larval and adult
mosquito Aedes aegypti. J. Exp. Biol. 209, 46384651.
Pelton, J.Z., 1938. The alimentary canal of the aphid Prociphilus Tesselata Fitch. Ohio J. Sci.
38, 164169.
Petzold-Maxwell, J.L., Cibile-Stewart, X., French, B.W., Gassmann, A.J., 2012. Adaptation
by western corn rootwrom (Coleoptera: Chrysomelidae) to Bt maize: inheritance, fitness
costs, and feeding preference. J. Econ. Entomol. 105, 14071418.
Ramsay, J.A., 1976. The rectal complex in the larvae of Lepidoptera. Philos. Trans. R. Soc.
Lond. B Biol. Sci. 274, 203226.
Ray, K., Mercedes, M., Chan, D., Choi, C.Y., Nishiura, J.T., 2009. Growth and differen-
tiation of the larval mosquito midgut. J. Insect Sci. 9, 5568.
Rheault, M.R., Okech, B.A., Keen, S.B.W., Miller, M.M., Meleshkevitch, E.A., Linser, P.J.,
Boubko, D.Y., Harvey, W.R., 2007. Molecular cloning, phylogeny and localization of
AgNHA1: the first Na+/H+ antiporter (NHA) from a metazoan, Anopheles gambiae.
J. Exp. Biol. 210, 38483861.
Rudin, W., Hecker, H., 1989. Lectin-binding sites in the midgut of the mosquitoes Anopheles
stephensi Liston and Aedes aegypti L. (Diptera: Culicidae). Parasitol. Res. 75, 268279.
Ryerse, J.S., Purcell, J.P., Sammons, R.D., 1994. Structure and formation of the peritrophic
membrane in the larva of the southern corn rootworm, Diabrotica undecimpunctata. Tissue
Cell 26, 431437.
Sayed, A., Nekl, E.R., Siqueira, A.A., Wang, H.C., Ffrenc-Constant, R.H., Bagley, M.,
Siegfried, B.D., 2007. A novel cadherin-like gene from western corn rootworm,
Diabrotica virgifera virgifera (Coleoptera: Chrysomelidae), larval midgut tissue. Insect
Mol. Biol. 16, 591600.
Seron, T.J., Hill, J., Linser, P.J., 2004. A GPI-linked carbonic anhydrase in the larval mos-
quito midgut. J. Exp. Biol. 207, 45594572.
Smith, K.E., VanEkeris, L.A., Linser, P.J., 2007. Cloning and characterization of AgCA9, a
novel -carbonic anhydrase from Anopheles gambiae Giles sensu stricto larvae. J. Exp. Biol.
210, 39193930.
Smith, K.E., VanEkeris, L.A., Okech, B.A., Harvey, W.R., Linser, P.J., 2008. Larval
anophleine mosquito recta exhibit a dramatic change in localization patterns of ion trans-
port proteins in response to shifting salinity: a comparison between anopheline and
culicine larvae. J. Exp. Biol. 211, 30673076.
Smith, K.E., VanEkeris, L., Valenti, M., Raymond, S., Smith, P., Linser, P.J., 2010. Phys-
iological and pharmacological characterization of the anopheline rectum shows redistri-
bution in function in response to varying salinity. Comp. Biochem. Physiol. C
157, 5562.
Terra, W.R., 1990. Evolution of digestive systems of insects. Annu. Rev. Entomol.
35, 181200.
Terra, W.R., Ferreira, C., Baker, J.E., 1996. Compartmentalization of digestion.
In: Lehane, M.J., Billingsley, P.F. (Eds.), Biology of the Insect Midgut. Chapman &
Hall, London, pp. 206235.
Insect Gut Structure, Function, Development and Target of Biological Toxins 37
Volkman, A., Peters, W., 1989a. Investigations on the midgut caeca of mosquito larvae-1.
Fine structure. Tissue Cell 21, 243251.
Volkman, A., Peters, W., 1989b. Investigations on the midgut caeca of mosquito larvae-II.
Functional aspects. Tissue Cell 21, 253261.
White, B.J., Kundert, P.N., Turissini, D.A., VanEkeris, L., Linser, P.J., Besansky, N.J., 2013.
Dose and developmental responses of Anopheles merus larvae to salinity. J. Exp. Biol.
216, 34333441.
Wieczorek, H., 1992. The insect V-ATPase, a plasma membrane proton pump energizing
secondary active transport: molecular analysis of electrogenic potassium transport in the
tobacco hornworm midgut. J. Exp. Biol. 172, 335343.
Xiang, M.A., Linser, P.J., Price, D.A., Harvey, W.R., 2012. Localization of two Na+- or
K+H+ antiporters, AgNHA1 and AgNHA2, in Anopheles gambiae larval Malpighian
tubules and the functional expression of AgNHA2 in yeast. J. Insect Physiol.
58, 570579.
Zhang, R., Hua, G., Andacht, T.M., Adang, M.J., 2008. A 106-kDa aminopeptidase is a
putative receptor for Bacillus thuringiensis Cry11Ba toxin in the mosquito Anopheles
gambiae. Biochemistry 47, 1126311272.
Zhuang, Z., Linser, P.J., Harvey, W.R., 1999. Antibody to H+ V-ATPase subunit
E colocalizes with portasomes in alkaline larval midgut of a freshwater mosquito (Aedes
aegypti). J. Exp. Biol. 202, 24492460.
CHAPTER TWO
Department of Biochemistry and Molecular Biology, University of Georgia, Athens, Georgia, USA
{
School of Life Sciences, University of Sussex, Falmer, Brighton, United Kingdom
}
Department of Entomology and Plant Pathology, University of Tennessee, Knoxville, Tennessee, USA
Contents
1. Introduction 40
2. General Characteristics of B. thuringiensis Crystal Toxins 42
2.1 Definition and classification of crystal toxins 42
2.2 The diversity of Cry toxins 43
2.3 The parasporin toxins 47
2.4 The ricin domain 48
2.5 Toxin discovery 49
3. Cry Toxin Structure: Function 49
3.1 Overview of Cry structure 49
3.2 Cry domain I 50
3.3 Cry domain II 51
3.4 Cry domain III 53
3.5 Cry intoxication process 55
3.6 Cry toxin solubilization and proteolytic processing 56
4. Midgut Cry-Binding Proteins and Receptor Function 57
4.1 Aminopeptidase 58
4.2 Cadherin 59
4.3 Alkaline phosphatase 60
4.4 ABC transporter 61
4.5 Other Cry-binding (receptor) proteins and molecules 62
5. Models of Cry Toxin Action 63
6. Cytolytic Toxins 68
Acknowledgements 70
References 70
Abstract
Parasporal crystals produced by Bacillus thuringiensis (Bt) bacteria are the main virulence
factors underlying Bt toxicity to insects. Parasporal crystals are composed primarily of Cry
and Cyt proteins that act on the midgut of susceptible insects. Cry proteins are an
important component of Bt biopesticides and are vital tools for insect control via expres-
sion in transgenic crop plants. Some members of the Cry group are more distantly
related including ETX/MTX and binary type toxins. Cry toxin structure and action
involves critical steps in toxin activation, binding to receptors such as cadherin and then
aminopeptidase or alkaline phosphatase probably in a sequential binding manner.
Specific Cry toxinreceptor interactions are a focus of this review. Recently, the impor-
tance of midgut ATP-binding cassette proteins to Cry intoxication of insects has been
demonstrated. Mechanistic details involved in sequential binding and pore formation
models are examined. The Cyt toxin of Bt subspecies israelensis is an important and inter-
esting component in Crymidgut interactions in mosquitoes. For some Cry toxins, Cyt
serves as a receptor for docking to midgut membrane. Recent engineering work has
demonstrated that Cyt can be re-targeted generating novel toxins for insect control.
Overall, we review the remarkable progress made in the past 20 years in discovering
novel Cry toxins and in elucidating complex mechanisms of Cry and Cyt toxin action;
subjects relevant to the long-term control of insects that damage crops and vector
human disease.
1. INTRODUCTION
The Gram-positive bacterium Bacillus thuringiensis (Bt) is characterized
by the proteinaceous crystals that it synthesizes in the mother cell during
sporulation (Aronson et al., 1986; Bulla et al., 1980). The history of the dis-
covery and development of Bt has been extensively reviewed (Beegle and
Yamamoto, 1992; Burges, 2001; Jurat-Fuentes and Jackson, 2012). Tens
of thousands of Bt strains have been isolated and most Bt strains are active
against larval stages of insects. Products based on Bt have been registered
as pesticides in the United States since 1961. As with other biological pes-
ticides, Bt offers a number of advantages over synthetic pesticides, including
lack of polluting residues, high specificity to target insects and safety to non-
target organisms. Consequently, the broadest uses of Bt are on food crops
and in forestry where safety and specific action are desirable. Disadvantages
of Bt are its high specificity and low persistence. While Bt is the most suc-
cessful biopesticide for insect control it remains a small part, about 2%, of the
total insecticide market. The widest usage of Bt for insect control is through
Diversity and Mechanisms of Bt Crystal Toxin Action 41
Clostridium bifermentans (Barloy et al., 1996) and Cry1Ia that is secreted rather
than being part of a crystal (Varani et al., 2013). In both these cases, the pro-
teins share significant sequence similarity to existing toxins and so are classified
as Cry toxins. For proteins that are found in the crystal, but share only weak
similarity to existing toxins, more evidence is required in order to classify
them. Normally, this would be a demonstration that either the purified toxin
or a recombinant form of the toxin has toxic activity to some target organism
or cell. In the case of a recombinant toxin, the protein would not actually have
to be produced by the parent strain. Indeed commercially important toxins,
such as Cry2Ab, are often encoded by cryptic genes (Crickmore et al., 1994).
The mere presence of a protein in the crystal of a B. thuringiensis strain is not
sufficient for that protein to be classified as a Cry toxin. The same criteria are
used when classifying Cyt toxins.
Figure 2.3 Venn diagram showing structural or functional relationships between toxins.
The individually coloured circles represent the different structural groups identified
in Fig. 2.2. Overlapping open blue circles indicate sequence similarity with non-Bt toxins
or proteins. The open red circles indicate which of the structural groups contain mem-
bers that share that particular characteristic namely binary toxin, parasporin or ricin
domain containing.
toxin in this group Cry46 (Parasporin 2), as mentioned above, does not
resemble any other Cry toxin but does have some similarity to hydralysin.
Although Parasporin 1 appears to be a typical three-domain toxin, its
mechanism of action is believed to be more complex than simple pore for-
mation (Katayama et al., 2007). In particular, the toxin induces an increase
in intracellular Ca2+ which is then associated with apoptosis as assessed by
increases in Caspase-3 and PARP cleavage. Less is known about the two
other three-domain toxins (Parasporins 3 and 6), although the cell swelling
observed after toxin administration is consistent with pore formation
(Nagamatsu et al., 2010; Yamashita et al., 2005). Parasporin 2 (Cry46) does
not share significant sequence similarity with other known Cry toxins but
its structure (Akiba et al., 2009) suggests that it ought to belong to the
ETX/MTX family. Its mechanism of action is believed to be through pore
formation after the toxin is targeted to lipid rafts on the cell surface where it
interacts with glycosylphosphatidyl inositol (GPI)-anchored proteins
(Kitada et al., 2009). Parasporin 4 appears to act as a cholesterol-
independent -pore-forming toxin that does not induce apoptosis
(Okumura et al., 2011). Nothing is currently known about the mechanism
of action of Parasporin 5.
Figure 2.4 Views of Cry1Aa toxins from the Resource for Structural Bioinformatics Protein
Data Bank (http://www.rcsb.org/pdb/) of (id: 1CIY) (Grochulski et al., 1995) modified using
PyMOL Version 1.7 (http://www.pymol.org/). Domain amino acid sequence position infor-
mation was obtained from NCBI data base (GenBank: AAP40639.1) and (Grochulski et al.,
1995). Overall views of Cry1Aa toxin are shown (Panels A and B). Domain I is red, domain
II is pink and domain III is orange. Domain I with the names of -helices and domain II with
named -helices and loops are shown in Panels C and D, respectively.
et al., 2009a), and nematicidal Cry5B toxin. Figure 2.4 shows views of the
Cry1Aa molecule generated from three-dimensional structure data
(Grochulski et al., 1995). The structure of domain I is a bundle of 78
-helices with a centrally located hydrophobic -helix 5. Based on observa-
tions that most of the -helices of domain I are long enough to span a hydro-
phobic cellular membrane and that domain I has similarities to pore-forming
domains in other bacterial toxins, it was hypothesized and later proven correct
that domain I is involved in membrane insertion and pore formation (Li et al.,
1991). Domain II, a three -sheet structure, is involved in receptor binding,
oligomerization, and membrane insertion. Domain III participates in receptor
binding and possibly membrane insertion.
mutagenizing residues S503 and S504 and attributed reduced M. sexta and
tobacco budworm, Heliothis virescens toxicity as being due to reduced initial
binding. How Cry1Ac binds molecules is unique among characterized
toxins, as the loop forms a lectin-like pocket and binds carbohydrates
(Burton et al., 1999; Jenkins et al., 1999). More specifically, this Cry1Ac
loop region participates in recognition of M. sexta midgut and in binding
to N-acetylgalactosamine (GalNAc) attached to aminopeptidase
(de Maagd et al., 1999b). It is also notable that the pattern of Cry1Ac
GalNAc recognition applies to the midgut ALPs of H. virescens and cotton
bollworm, Helicoverpa armigera ( Jurat-Fuentes and Adang, 2004; Sarkar et al.,
2009; Sengupta et al., 2013) and that mutations in the GalNAc binding
region reduced toxicity. Recently, the GalNAc binding pocket on Cry1Ac
was disrupted by amino acid replacements resulting in reduced binding to
monomeric ALP and insecticidal activity. The results of Pardo-Lopez
et al. (2006) indicate that GalNAc binding to Cry1Ac induces a slight con-
formational change that enhances membrane insertion of an oligomeric pre-
pore structure. The relevance of Cry1Ac domain III discussed above to
binding and toxicity is also buttressed by studies showing that domain III
of Cry1Ac mediates specificity in lepidopteran species including L. dispar
(Lee et al., 1995), H. virescens and the cabbage looper, Trichoplusia ni
(Ge et al., 1991). Construction of Cry1Cry1Ac hybrid toxins where
Cry1Ac is attached C-terminal to domains III from another Cry1 toxin
in some cases resulted in Cry toxins more toxic to H. virescens (Karlova
et al., 2005). Studies of Cry1C, an important toxin for its activity against
armyworm, Spodoptera species, identified domain III as a critical determinant
in insecticidal specificity. Similar to the Cry1Ac domain III loop discussed
above, Cry1C has a predicted loop connecting two -strands on the outer
-sheet (Herrero et al., 2004). The importance of this region in Cry1C was
suggested by hybrid toxin G27 (domains I and II from Cry1Ea and domain
III from Cry1Ca) when two residues were substituted and toxicity to
Spodoptera exigua, but not M. sexta larvae was reduced (de Maagd et al.,
1999a). Recently, Herrero et al. (2004) showed that mutation of residues
in the Cry1Ca domain III loop reduced toxicity to beet armyworm,
S. exigua, larvae and binding to BBMV, but did not alter oligomerization.
Transfer of toxicity against Spodoptera species from Cry1C to Cry1Ac was
achieved by transfer of Cry1C domain III (Bosch et al., 1994). A novel
domain III exchange involved the transfer of Cry1Ab domain III onto
Cry3Aa domains III in coleopteran activity that did not compete for bind-
ing with Cry3Aa-like proteins (Walters et al., 2010). This result supports an
Diversity and Mechanisms of Bt Crystal Toxin Action 55
important binding role for the lepidopteran-active Cry1Ab domain III that
is functional in the coleopteran western corn rootworm, Diabrotica virgifera.
Figure 2.5 Representation of the current models of Cry toxin action in the insect midgut
epithelium. After ingestion by susceptible larvae, toxin crystals are solubilized in the mid-
gut fluids to yield the Cry protoxin form, which is processed to an activated Cry toxin form
equivalent to the Cry toxin ingested by insects feeding on transgenic Bt crops. The Cry
toxin core then traverses the peritrophic matrix, which is able to retain some of the toxin
(darker toxin molecules in the figure). Once reaching the brush border membrane of the
midgut epithelium, the Cry toxin binds with high affinity to cadherin, which results in acti-
vation of intracellular cell death pathways (represented by G protein, adenylate cyclase
and cAMP in the figure), and/or according to the sequential binding model, further pro-
teolysis of the toxin monomer to result in formation of a pre-pore oligomer. This pre-pore
oligomer is proposed to bind to alkaline phosphatase (ALP) or aminopeptidase (APN) to
insert in the membrane forming a pore that leads to osmotic cell death. The insertion of
toxin monomer and formation of a pore by oligomerization of inserted monomers is also
presented as an alternative step. Disruption of the midgut epithelium barrier allows for
bacterial invasion of the haemocoel, leading to septicaemia and death of the insect.
56 Michael J. Adang et al.
4.1. Aminopeptidase
Aminopeptidase-N (APN) (Knight et al., 1994, 1995; Sangadala et al., 1994)
and cadherin (Vadlamudi et al., 1995) were the first proteins identified as
putative Cry receptors in insects. Having two distinct proteins types as puta-
tive receptor raised questions as to whether or not both proteins were
involved in Cry toxin action; an issue that is resolved in some of the pro-
posed intoxication models, as discussed below. APNs are tethered to the
midgut brush border by GPI anchors (Garczynski and Adang, 1995) from
where they cleave N-terminal amino acids from peptides, a step necessary
for amino acid co-transport into epithelial cells (Terra et al., 1996). Phylo-
genetic analyses of lepidopteran APNs cluster these proteins into seven clas-
ses (Crava et al., 2010; Hughes, 2014). Among APNs for which expression
data are available, members of the APN1 class are the most highly expressed
in midgut tissue (Hughes, 2014). Interactions between Cry1 toxins and mid-
gut APNs are known to involve recognition of epitopes on the primary pro-
tein structure, or in the case of Cry1Ac, an attached glycan with a terminal
GalNAc moiety. APNs that bind Cry1Ac via GalNAc are identified in
M. sexta (Masson et al., 1995), H. virescens (Gill et al., 1995; Luo et al.,
1997), L. dispar ( Jenkins et al., 2000) and H. armigera (Sarkar et al., 2009).
Cry1Ac binding to APNs in M. sexta and L. dispar has sugar-dependent
and -independent components ( Jenkins et al., 2000; Masson et al., 1995)
leading to proposal of a bivalent binding model involving sequential inter-
actions with domain III and then domain II loop residues ( Jenkins et al.,
2000). Cry1Ac also recognizes a 106-kDa APN in H. virescens in a
GalNAc-independent manner (Banks et al., 2001). Similarly, and although
only known to recognize receptor proteins at amino acid epitopes, the
Cry1Aa toxin binds through domains II and III to a 117-kDa APN in
B. mori (Atsumi et al., 2005). A small patch of seven amino acids near the
N-terminus of a cotton leafworm, Spodoptera litura APN serves as epitope
for recognition by loops 2 and 3 of Cry1C domain II (Kauer et al., 2014).
Diversity and Mechanisms of Bt Crystal Toxin Action 59
In the 1990s, evidence for APN function as Cry1 receptors was provided
by the ability of an APN preparation to enhance Cry1-induced pore forma-
tion in membrane vesicles (Luo et al., 1997; Sangadala et al., 1994) and ion
channels in membrane bilayers (Schwartz et al., 1997c). This approach was
particularly challenging due to the difficulty of purifying midgut APNs with
the lipid moiety remaining on the GPI anchor (Garczynski and Adang,
1995). In vivo evidence of Cry receptor was suggested by work of Gill
and Ellar (2002) who expressed M. sexta APN in transgenic Drosophila ren-
dering larvae susceptible to Cry1Ac. Silencing midgut APNs in S. litura, sug-
arcane borer (Diatraea saccharalis) and H. armigera by RNA interference
(RNAi) established their role as Cry1C (Rajagopal et al., 2002), Cry1Ab
(Yang et al., 2010), and Cry1Ac (Sivakumar et al., 2007) receptors, respec-
tively. Mutation or reduced expression of specific APNs has been correlated
with resistance to Cry1 toxins in D. saccharalis (Yang et al., 2010), S. exigua
(Herrero et al., 2005) and H. armigera (Zhang et al., 2009). In a greenhouse-
derived strain of T. ni, Cry1Ac resistance was associated with differential
alteration of two APNs; with APN1 being downregulated and APN6 being
upregulated in resistant larvae (Tiewsiri and Wang, 2011). Also, the lack of
110-kDa APN1 protein in resistant larvae correlated with reduced transcript
levels and was conferred by a trans-regulatory mechanism (Tiewsiri and
Wang, 2011).
The role of APNs in the action of mosquitocidal Cry toxins has received
considerable attention. Diverse APNs have been identified as putative
receptors of Cry11Aa and Cry11Ba in yellow fever mosquito, Aedes aegypti
(Chen et al., 2009b, 2013; Likitvivatanavong et al., 2011), and Cry11Ba in
malaria mosquitoes, Anopheles albimanus and An. gambiae (Abdullah et al.,
2006; Zhang et al., 2008). The specific APNs in Aedes and Anopheles that
bind Cry11Aa and Cry11Ba, respectively, do so with high affinity (nM
range). Interestingly, in bioassays, the presence of partial APN fragments
enhanced Cry larval mortality (Chen et al., 2013; Zhang et al., 2010), an
unexpected observation from a functional receptor that may suggest inter-
actions between APN and toxin are reversible. Silencing of APN expression
in Ae. aegypti by RNAi resulted in increased tolerance to Cry4Ba toxin
(Saengwiman et al., 2011), supporting a functional toxin receptor role for
this APN.
4.2. Cadherin
Cadherin-like proteins are widely accepted as functional Cry toxin recep-
tors. Contrary to the generally observed localization for cadherin proteins
60 Michael J. Adang et al.
domain I insert into the membrane, while the rest of the domain lays flat on
the membrane surface. Alternatively, most available evidence supports the
umbrella model, in which a conformational change in the toxin molecule
upon binding to receptors results in insertion of a hydrophobic hairpin com-
posed of 4 and 5 helices (Li et al., 2001; Schwartz et al., 1997c). The posi-
tion of domains II and III of the Cry toxins during insertion is still a matter of
debate. While protease protection studies (Aronson, 2000; Aronson et al.,
1999) and mutagenesis (Dean et al., 1996; Nair and Dean, 2008) support that
the whole toxin molecule is confined to the membrane, there is also evi-
dence suggesting that the toxin remains associated to membrane receptors
after pore formation (Fortier et al., 2007) and at least residues of domain
III may be exposed to the solvent (Pardo-Lopez et al., 2006). In contrast
to the pre-pore oligomers presented by the sequential binding model, pro-
posals modelling pore formation support that association between inserted
toxin monomers results in a tetrameric ion channel with a diameter of
approximately 15 A (Groulx et al., 2010, 2011). However, trimeric pores
have been proposed from sequence analysis (Torres et al., 2008) and lipid
membrane experiments (Ounjai et al., 2007). Evidence supporting the pos-
sibility that Cry toxin pores may arrange from association of diverse Cry
toxin monomers (heterooligomers) has been presented (Carmona et al.,
2011), which may help explain synergistic and inhibitory effects among
Cry toxins (Ibargutxi et al., 2008). More importantly, the significance of
heterooligomeric pore formation for binding competition studies needs
to be considered and tested experimentally.
The 4 helix lining the Cry toxin pore seems to control the passage of
ions through the pore (Kumar and Aronson, 1999; Masson et al., 1999),
although there are also reports suggesting a role in controlling pore perme-
ability by alternative domain I helices (Alcantara et al., 2001; Arnold et al.,
2001). Pore properties as well as formation are also dependent on midgut
epithelium components (Peyronnet et al., 2001; Schwartz et al., 1997b)
and ionic composition (Brunet et al., 2010a; Fortier et al., 2005). Variable
pore properties depending on the environment may reflect adaptation of
Cry toxins to diverse functional environments (Schnepf et al., 1998).
An alternative model for Cry intoxication disregards pore formation and
highlights the activation of intracellular oncotic cell death pathways as
responsible for enterocyte death (Zhang et al., 2008). However, evidence
supporting this model is limited to Cry1Ab and studies with insect cell cul-
tures expressing a cadherin from M. sexta (Zhang et al., 2005). In these cells,
binding of Cry1Ab toxin was associated with cell death and increased cAMP
Diversity and Mechanisms of Bt Crystal Toxin Action 67
production. It is important to note that the ovarian Hi5 cell line used for
these analyses is susceptible to the Cry1Ac toxin in the absence of expression
of proposed midgut toxin receptors, suggesting that they may display a
mechanism of susceptibility that differs from cells in the insect midgut epi-
thelium. Evidence supporting a role for intracellular pathways in Cry intox-
ication of midgut cells was provided by a deletion of 55 amino acids in the
cytoplasmic domain of a cadherin being genetically linked with non-
recessive resistance to Cry1Ac in H. armigera (Zhang et al., 2012). However,
only a 20% reduction in susceptibility to Cry1Ac was detected in Sf9 cells
expressing the wild type versus the cadherin with the cytoplasmic deletion,
challenging the relative importance of cadherin-mediated intracellular sig-
nalling in Cry susceptibility.
Although speculative, a third possibility would consider effects of both
pore formation and intracellular cell death pathways (Pigott and Ellar,
2007). Thus, there are examples of bacterial pore-forming toxins that induce
host cell death through oncosis (Zhou et al., 2009). The observation that
modified Cry toxins that do not depend on binding to cadherin for oligo-
merization display lower activity against susceptible insects (Soberon et al.,
2007), may suggest that intracellular signalling activated by wild type, but
not by modified Cry toxins binding to cadherin may contribute to toxicity.
Consequently, it is plausible that intracellular cell death may occur in the
presence of lower Cry toxin concentrations, while higher toxin concentra-
tions would promote increased toxin insertion and formation of pores. Fur-
ther work to establish functional connections between pore formation and
intracellular signalling for cytotoxicity are needed to clarify the molecular
events resulting in midgut cell death by Cry toxins.
In the case of binary Cry toxins, low sequence homology suggests dif-
ferences in the intoxication model when compared with the three-domain
Cry toxins. Most data available are focused on the Cry34/35 complex (Ellis
et al., 2002). In this case, while the Cry35 protein displays lectin folds sug-
gestive of receptor binding and the Cry34 protein has homology to pro-
teins involved in intracellular signalling (Schnepf et al., 2005), they
probably exert their toxicity through pore formation (Masson et al.,
2004). Data from binding assays support that Cry34 toxin greatly enhances
Cry35 binding to D. virgifera BBMV proteins (Li et al., 2013). Interestingly,
comparisons between results from homologous competition tests in the
presence or absence of Cry34 suggest lower binding affinity for the
Cry34/35 complex. Binding of Cry35 to D. virgifera BBMV proteins in
the absence or presence of Cry34 was to binding sites not recognized
68 Michael J. Adang et al.
by Cry3Aa, Cry8Ba or Cry6Aa toxins (Li et al., 2013). The identity of the
specific Cry34/35 or Cry35 receptors is unknown.
6. CYTOLYTIC TOXINS
The Cyt (cytolytic) toxins were originally discovered in Bt subspecies
israelensis (Goldberg and Margalit, 1977) and have been recently reviewed
(Ben-Dov, 2014; Soberon et al., 2013b). These proteins seem to be specific
to some Bt subspecies (Tyrell et al., 1981). The first feature observed for the
28 kDa protein from Bt subsp. israelensis was its cytolytic activity against cul-
tured mammalian and insects cells and toxicity when injected in mice, while
it did not display activity against larvae of the cabbage butterfly, Pieris brassicae
(Thomas and Ellar, 1983a). Against mosquitoes, the 28-kDa protein was less
active than native Bt subsp. israelensis crystals (Chilcott and Ellar, 1988;
Davidson and Yamamoto, 1984; Yamamoto et al., 1983), and it synergized
activity of other Bt subsp. israelensis crystal proteins (Chilcott and Ellar, 1988;
Wu and Chang, 1985). Currently, there are also cyt genes that have been
described in Bt strains targeting lepidopteran or coleopteran insects
(Guerchicoff et al., 1997), adding to, three cyt toxin gene families (cyt1,
cyt2 and cyt3) that include 11 holotype toxins in the current nomenclature
(Crickmore et al., 2014).
The three-dimensional Cyt structures resolved to date support a con-
served structural model including two -helix hairpins flanking a
-sheet core containing seven to eight -strands (Cohen et al., 2008,
2011; Li et al., 1996). This highly conserved structure is also revealed in
sequence alignments, which identify highly conserved blocks (Butko,
2003). Mutagenic studies identified -sheet residues as critical for toxicity,
while mutations of residues on the helical domains did not affect toxicity,
suggesting a critical role for the -sheet core.
Ingested Cyt1A and Cyt2A protoxins are processed by digestive prote-
ases at the same sites in the N- and C-termini to a stable toxin core of 25 and
23 kDa, respectively (Koni and Ellar, 1994). Activated Cyt toxins display
high affinity for membrane lipids containing unsaturated acyl chains (Gill
et al., 1987; Thomas and Ellar, 1983b), which are abundant in midgut brush
border membrane of Diptera (Li et al., 1996). A putative phospholipid-
binding site pocket homologous to Erwinia virulence factor (Evf ) was
described for Cyt2Ba (Rigden, 2009). Two non-exclusive mechanisms
for Cyt1A-induced cytolysis have been proposed: pore formation and
detergent-like membrane disruption (Butko, 2003). Pore formation by
Diversity and Mechanisms of Bt Crystal Toxin Action 69
ACKNOWLEDGEMENTS
This work was supported by USDA NIFA award number 2010-65105-20590 to M. J. A.
(University of Georgia) and Biotechnology Risk Assessment Grant Program competitive
grant no. 2010-33522-21700 from USDA NIFA to J. L. J-F. (University of Tennessee).
We thank Dr. Leara Rhodes (University of Georgia) for editing an earlier version of this
manuscript. Ruchir Mishra is thanked for generating and assembling the three-
dimensional Cry images.
REFERENCES
Abdullah, M.A., Valaitis, A.P., Dean, D.H., 2006. Identification of a Bacillus thuringiensis
Cry11Ba toxin-binding aminopeptidase from the mosquito, Anopheles quadrimaculatus.
BMC Biochem. 7, 16.
Ai, B., Li, J., Feng, D., Li, F., Guo, S., 2013. The elimination of DNA from the Cry toxin-
DNA complex is a necessary step in the mode of action of the Cry8 toxin. PLoS One 8.
Aimanova, K.G., Zhuang, M., Gill, S.S., 2006. Expression of Cry1Ac cadherin receptors in
insect midgut and cell lines. J. Invertebr. Pathol. 92, 178187.
Akiba, T., Abe, Y., Kitada, S., Kusaka, Y., Ito, A., Ichimatsu, T., Katayama, H., Akao, T.,
Higuchi, K., Mizuki, E., Ohba, M., Kanai, R., Harata, K., 2009. Crystal structure of the
parasporin-2 Bacillus thuringiensis toxin that recognizes cancer cells. J. Mol. Biol.
386, 121133.
Alcantara, E.P., Alzate, O., Lee, M.K., Curtiss, A., Dean, D.H., 2001. Role of alpha-helix
seven of Bacillus thuringiensis Cry1Ab delta-endotoxin in membrane insertion, structural
stability, and ion channel activity. Biochemistry 40, 25402547.
Alzate, O., Osorio, C., Florez, A.M., Dean, D.H., 2010. Participation of valine 171 in alpha-
Helix 5 of Bacillus thuringiensis Cry1Ab delta-endotoxin in translocation of toxin into
Lymantria dispar midgut membranes. Appl. Environ. Microbiol. 76, 78787880.
Andrews Jr., R.E., Bibilos, M.M., Bulla Jr., L.A., 1985. Protease activation of the
entomocidal protoxin of Bacillus thuringiensis ssp. kurstaki. Appl. Environ. Microbiol.
50, 737742.
Angelucci, C., Barrett-Wilt, G.A., Hunt, D.F., Akhurst, R.J., East, P.D., Gordon, K.H.,
Campbell, P.M., 2008. Diversity of aminopeptidases, derived from four lepidopteran
gene duplications, and polycalins expressed in the midgut of Helicoverpa armigera: iden-
tification of proteins binding the delta-endotoxin, Cry1Ac of Bacillus thuringiensis. Insect
Biochem. Mol. Biol. 38, 685696.
Arenas, I., Bravo, A., Soberon, M., Gomez, I., 2010. Role of alkaline phosphatase from Man-
duca sexta in the mechanism of action of Bacillus thuringiensis Cry1Ab toxin. J. Biol.
Chem. 285, 1249712503.
Arnold, S., Curtiss, A., Dean, D.H., Alzate, O., 2001. The role of a proline-induced broken-
helix motif in alpha-helix 2 of Bacillus thuringiensis delta-endotoxins. FEBS Lett.
490, 7074.
Diversity and Mechanisms of Bt Crystal Toxin Action 71
Aronson, A., 2000. Incorporation of protease K into larval insect membrane vesicles does not
result in disruption of integrity or function of the pore-forming Bacillus thuringiensis delta-
endotoxin. Appl. Environ. Microbiol. 66, 45684570.
Aronson, A.I., Beckman, W., Dunn, P., 1986. Bacillus thuringiensis and related insect path-
ogens. Microbiol. Rev. 50, 124.
Aronson, A.I., Geng, C., Wu, L., 1999. Aggregation of Bacillus thuringiensis Cry1A toxins
upon binding to target insect larval midgut vesicles. Appl. Environ. Microbiol.
65, 25032507.
Aronson, A.I., Han, E.S., McGaughey, W., Johnson, D., 1991. The solubility of inclusion
proteins from Bacillus thuringiensis is dependent upon protoxin composition and is a factor
in toxicity to insects. Appl. Environ. Microbiol. 57, 981986.
Aronson, A.I., Wu, D., Zhang, C., 1995. Mutagenesis of specificity and toxicity regions of a
Bacillus thuringiensis protoxin gene. J. Bacteriol. 177, 40594065.
Atsumi, S., Miyamoto, K., Yamamoto, K., Narukawa, J., Kawai, S., Sezutsu, H.,
Kobayashi, I., Uchino, K., Tamura, T., Mita, K., Kadono-Okuda, K., Wada, S.,
Kanda, K., Goldsmith, M.R., Noda, H., 2012. Single amino acid mutation in an
ATP-binding cassette transporter gene causes resistance to Bt toxin Cry1Ab in the silk-
worm, Bombyx mori. Proc. Natl. Acad. Sci. U.S.A. 109, E1591E1598.
Atsumi, S., Mizuno, E., Hara, H., Nakanishi, K., Kitami, M., Miura, N., Tabunoki, H.,
Watanabe, A., Sato, R., 2005. Location of the Bombyx mori aminopeptidase N type 1 binding
site on Bacillus thuringiensis Cry1Aa toxin. Appl. Environ. Microbiol. 71, 39663977.
Banks, D.J., Jurat-Fuentes, J.L., Dean, D.H., Adang, M.J., 2001. Bacillus thuringiensis Cry1Ac
and Cry1Fa delta-endotoxin binding to a novel 110 kDa aminopeptidase in Heliothis vir-
escens is not N-acetylgalactosamine mediated. Insect Biochem. Mol. Biol. 31, 909918.
Barloy, F., Delecluse, A., Nicolas, L., Lecadet, M.M., 1996. Cloning and expression of the
first anaerobic toxin gene from Clostridium bifermentans subsp. malaysia, encoding a new
mosquitocidal protein with homologies to Bacillus thuringiensis delta-endotoxins.
J. Bacteriol. 178, 30993105.
Baxter, S.W., Badenes-Perez, F.R., Morrison, A., Vogel, H., Crickmore, N., Kain, W.,
Wang, P., Heckel, D.G., Jiggins, C.D., 2011. Parallel evolution of Bacillus thuringiensis
toxin resistance in Lepidoptera. Genetics 189, 675679.
Bayyareddy, K., Andacht, T.M., Abdullah, M.A., Adang, M.J., 2009. Proteomic identifica-
tion of Bacillus thuringiensis subsp. israelensis toxin Cry4Ba binding proteins in midgut
membranes from Aedes (Stegomyia) aegypti Linnaeus (Diptera, Culicidae) larvae. Insect
Biochem. Mol. Biol. 39, 279286.
Beegle, C.C., Yamamoto, T., 1992. History of Bacillus thuringiensis Berliner research and
development. Can. Entomol. 124, 587616.
Ben-Dov, E., 2014. Bacillus thuringiensis subsp. israelensis and its dipteran-specific toxins.
Toxins (Basel) 6, 12221243.
Berne, S., Lah, L., Sepcic, K., 2009. Aegerolysins: structure, function, and putative biological
role. Protein Sci. 18, 694706.
Berry, C., 2012. The bacterium, Lysinibacillus sphaericus, as an insect pathogen. J. Invertebr.
Pathol. 109, 110.
Bietlot, H.P., Carey, P.R., Choma, C., Kaplan, H., Lessard, T., Pozsgay, M., 1989. Facile
preparation and characterization of the toxin from Bacillus thuringiensis var. kurstaki. Bio-
chem. J. 260, 8791.
Bietlot, H.P., Schernthaner, J.P., Milne, R.E., Clairmont, F.R., Bhella, R.S., Kaplan, H.,
1993. Evidence that the CryIA crystal protein from Bacillus thuringiensis is associated with
DNA. J. Biol. Chem. 268, 82408245.
Bokori-Brown, M., Savva, C.G., Fernandes da Costa, S.P., Naylor, C.E., Basak, A.K.,
Titball, R.W., 2011. Molecular basis of toxicity of Clostridium perfringens epsilon toxin.
FEBS J. 278, 45894601.
72 Michael J. Adang et al.
Bonin, A., Paris, M., Tetreau, G., David, J.P., Despres, L., 2009. Candidate genes revealed by
a genome scan for mosquito resistance to a bacterial insecticide: sequence and gene
expression variations. BMC Genomics 10, 551.
Boonserm, P., Davis, P., Ellar, D.J., Li, J., 2005. Crystal structure of the mosquito-larvicidal
toxin Cry4Ba and its biological implications. J. Mol. Biol. 348, 363382.
Boonserm, P., Mo, M., Angsuthanasombat, C., Lescar, J., 2006. Structure of the functional
form of the mosquito larvicidal Cry4Aa toxin from Bacillus thuringiensis at a 2.8-angstrom
resolution. J. Bacteriol. 188, 33913401.
Bosch, D., Schipper, B., van der Kleij, H., de Maagd, R.A., Stiekema, W.J., 1994. Recom-
binant Bacillus thuringiensis crystal proteins with new properties: possibilities for resistance
management. Nat. Biotechnol. 12, 915918.
Bravo, A., Gomez, I., Conde, J., Munoz-Garay, C., Sanchez, J., Miranda, R., Zhuang, M.,
Gill, S.S., Soberon, M., 2004. Oligomerization triggers binding of a Bacillus thuringiensis
Cry1Ab pore-forming toxin to aminopeptidase N receptor leading to insertion into
membrane microdomains. Biochim. Biophys. Acta 1667, 3846.
Bravo, A., Hendrickx, K., Jansens, S., Peferoen, M., 1992. Immunocytochemical analysis of
specific binding of Bacillus thuringiensis insecticidal crystal proteins to lepidopteran and
coleopteran midgut membranes. J. Invertebr. Pathol. 60, 247253.
Bravo, A., Likitvivatanavong, S., Gill, S.S., Soberon, M., 2011. Bacillus thuringiensis: a story of
a successful bioinsecticide. Insect Biochem. Mol. Biol. 41, 423431.
Bravo, A., Sanchez, J., Kouskoura, T., Crickmore, N., 2002. N-terminal activation is an
essential early step in the mechanism of action of the Bacillus thuringiensis Cry1Ac insec-
ticidal toxin. J. Biol. Chem. 277, 2398523987.
Brunet, J.F., Vachon, V., Juteau, M., Van Rie, J., Larouche, G., Vincent, C., Schwartz, J.L.,
Laprade, R., 2010a. Pore-forming properties of the Bacillus thuringiensis toxin Cry9Ca in
Manduca sexta brush border membrane vesicles. Biochim. Biophys. Acta Biomembr.
1798, 11111118.
Brunet, J.F., Vachon, V., Marsolais, M., Van Rie, J., Schwartz, J.L., Laprade, R., 2010b.
Midgut juice components affect pore formation by the Bacillus thuringiensis insecticidal
toxin Cry9Ca. J. Invertebr. Pathol. 104, 203208.
Bulla Jr., L.A., Bechtel, D.B., Kramer, K.J., Shethna, Y.I., Aronson, A.I., Fitz-James, P.C.,
1980. Ultrastructure, physiology, and biochemistry of Bacillus thuringiensis. Crit. Rev.
Microbiol. 8, 147204.
Burges, H.D., 2001. Bacillus thuringiensis in pest control. Pestic. Outlook 12, 9098.
Burton, S.L., Ellar, D.J., Li, J., Derbyshire, D.J., 1999. N-acetylgalactosamine on the putative
insect receptor aminopeptidase N is recognised by a site on the domain III lectin-like fold
of a Bacillus thuringiensis insecticidal toxin. J. Mol. Biol. 287, 10111022.
Butko, P., 2003. Cytolytic toxin Cyt1A and its mechanism of membrane damage: data and
hypotheses. Appl. Environ. Microbiol. 69, 24152422.
Butko, P., Huang, F., PusztaiCarey, M., Surewicz, W.K., 1996. Membrane permeabilization
induced by cytolytic delta-endotoxin CytA from Bacillus thuringiensis var israelensis.
Biochemistry 35, 1135511360.
Canton, P.E., Lopez-Daz, J.A., Gill, S.S., Bravo, A., Soberon, M., 2014. Membrane binding
and oligomer membrane insertion are necessary but insufficient for Bacillus thuringiensis
Cyt1Aa toxicity. Peptides 53, 286291.
Carmona, D., Rodrguez-Almazan, C., Munoz-Garay, C., Portugal, L., Perez, C., de
Maagd, R.A., Bakker, P., Soberon, M., Bravo, A., 2011. Dominant negative phenotype
of Bacillus thuringiensis Cry1Ab, Cry11Aa and Cry4Ba mutants suggest hetero-oligomer
formation among different Cry toxins. PLoS One 6.
Carroll, J., Convents, D., Van Damme, J., Boets, A., Van Rie, J., Ellar, D.J., 1997. Intramo-
lecular proteolytic cleavage of Bacillus thuringiensis Cry3A delta-endotoxin may facilitate
its coleopteran toxicity. J. Invertebr. Pathol. 70, 4149.
Diversity and Mechanisms of Bt Crystal Toxin Action 73
Chen, J., Aimanova, K.G., Fernandez, L.E., Bravo, A., Soberon, M., Gill, S.S., 2009a. Aedes
aegypti cadherin serves as a putative receptor of the Cry11Aa toxin from Bacillus thur-
ingiensis subsp. israelensis. Biochem. J. 424, 191200.
Chen, J., Aimanova, K.G., Pan, S., Gill, S.S., 2009b. Identification and characterization of
Aedes aegypti aminopeptidase N as a putative receptor of Bacillus thuringiensis Cry11A
toxin. Insect Biochem. Mol. Biol. 39, 688696.
Chen, J., Brown, M.R., Hua, G., Adang, M.J., 2005. Comparison of the localization of Bacil-
lus thuringiensis Cry1A delta-endotoxins and their binding proteins in larval midgut of
tobacco hornworm, Manduca sexta. Cell Tissue Res. 321, 123129.
Chen, J., Likitvivatanavong, S., Aimanova, K.G., Gill, S.S., 2013. A 104 kDa Aedes aegypti
aminopeptidase N is a putative receptor for the Cry11Aa toxin from Bacillus thuringiensis
subsp. israelensis. Insect Biochem. Mol. Biol. 43, 12011208.
Chen, X.J., Curtiss, A., Alcantara, E., Dean, D.H., 1995. Mutations in domain I of Bacillus
thuringiensis d-endotoxin CryIAb reduce the irreversible binding of toxin to Manduca
sexta brush border membrane vesicles. J. Biol. Chem. 270, 64126419.
Chen, X.J., Lee, M.K., Dean, D.H., 1993. Site-directed mutations in a highly conserved
region of Bacillus thuringiensis -endotoxin affect inhibition of short circuit current across
Bombyx mori midguts. Proc. Natl. Acad. Sci. U.S.A. 90, 90419045.
Chilcott, C.N., Ellar, D.J., 1988. Comparative toxicity of Bacillus thuringiensis var. israelensis
crystal proteins in-vivo and in-vitro. J. Gen. Microbiol. 134, 25512558.
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, 84658470.
Chow, E., Singh, G.J., Gill, S.S., 1989. Binding and aggregation of the 25-kilodalton toxin of
Bacillus thuringiensis subsp. israelensis to cell membranes and alteration by monoclonal anti-
bodies and amino acid modifiers. Appl. Environ. Microbiol. 55, 27792788.
Clairmont, F.R., Milne, R.E., Pham, V.T., Carriere, M.B., Kaplan, H., 1998. Role of DNA
in the activation of the Cry1A insecticidal crystal protein from Bacillus thuringiensis.
J. Biol. Chem. 273, 92929296.
Cohen, S., Albeck, S., Ben-Dov, E., Cahan, R., Firer, M., Zaritsky, A., Dym, O., 2011.
Cyt1Aa toxin: crystal structure reveals implications for its membrane-perforating func-
tion. J. Mol. Biol. 413, 804814.
Cohen, S., Dym, O., Albeck, S., Ben-Dov, E., Cahan, R., Firer, M., Zaritsky, A., 2008.
High-resolution crystal structure of activated Cyt2Ba monomer from Bacillus thuringiensis
subsp. israelensis. J. Mol. Biol. 380, 820827.
Crava, C.M., Bel, Y., Lee, S.F., Manachini, B., Heckel, D.G., Escriche, B., 2010. Study of the
aminopeptidase N gene family in the lepidopterans Ostrinia nubilalis (Hubner) and Bombyx
mori (L.): sequences, mapping and expression. Insect Biochem. Mol. Biol. 40, 506515.
Crickmore, N., Baum, J., Bravo, A., Lereclus, D., Narva, K., Sampson, K., Schnepf, E., Sun,
M., Zeigler, D.R., 2014. Bacillus thuringiensis toxin nomenclature. http://www.lifesci.
sussex.ac.uk/home/Neil_Crickmore/Bt/.
Crickmore, N., Wheeler, V.C., Ellar, D.J., 1994. Use of an operon fusion to induce expres-
sion and crystallisation of a Bacillus thuringiensis delta-endotoxin encoded by a cryptic
gene. Mol. Gen. Genet. 242, 365368.
Crickmore, N., Zeigler, D.R., Feitelson, J., Schnepf, E., Van Rie, J., Lereclus, D., Baum, J.,
Dean, D.H., 1998. Revision of the nomenclature for the Bacillus thuringiensis pesticidal
crystal proteins. Microbiol. Mol. Biol. Rev. 62, 807813.
Davidson, E.W., Yamamoto, T., 1984. Isolation and assay of the toxic component from the
crystals of Bacillus thuringiensis var. israelensis. Curr. Microbiol. 11, 171174.
de Maagd, R.A., Bakker, P., Staykov, N., Dukiandjiev, S., Stiekema, W., Bosch, D., 1999a.
Identification of Bacillus thuringiensis delta-endotoxin Cry1C domain III amino acid res-
idues involved in insect specificity. Appl. Environ. Microbiol. 65, 43694374.
74 Michael J. Adang et al.
de Maagd, R.A., Bakker, P.L., Masson, L., Adang, M.J., Sangadala, S., Stiekema, W.,
Bosch, D., 1999b. Domain III of the Bacillus thuringiensis delta-endotoxin Cry1Ac is
involved in binding to Manduca sexta brush border membranes and to its purified ami-
nopeptidase N. Mol. Microbiol. 31, 463471.
de Maagd, R.A., Bravo, A., Berry, C., Crickmore, N., Schnepf, H.E., 2003. Structure,
diversity, and evolution of protein toxins from spore-forming entomopathogenic bacte-
ria. Annu. Rev. Genet. 37, 409433.
de Maagd, R.A., Bravo, A., Crickmore, N., 2001. How Bacillus thuringiensis has evolved spe-
cific toxins to colonize the insect world. Trends Genet. 17, 193199.
Dean, D.H., Rajamohan, F., Lee, M.K., Wu, S.J., Chen, X.J., Alcantara, E., Hussain, S.R.,
1996. Probing the mechanism of action of Bacillus thuringiensis insecticidal proteins by
site-directed mutagenesisa mini review. Gene 179, 111117.
Dechklar, M., Tiewsiri, K., Angsuthanasombat, C., Pootanakit, K., 2011. Functional expres-
sion in insect cells of glycosylphosphatidylinositol-linked alkaline phosphatase from Aedes
aegypti larval midgut: a Bacillus thuringiensis Cry4Ba toxin receptor. Insect Biochem. Mol.
Biol. 41, 159166.
Dennis, R.D., Wiegandt, H., Haustein, D., Knowles, B., Ellar, D.J., 1986. Thin layer chro-
matography overlay technique in the analysis of the binding of the solubilized protoxin of
Bacillus thuringiensis var. kurstaki to an insect glycosphingolipid of known structure.
Biomed. Chromatogr. 1, 3137.
Dermauw, W., Van Leeuwen, T., 2014. The ABC gene family in arthropods: comparative
genomics and role in insecticide transport and resistance. Insect Biochem. Mol. Biol.
45, 89110.
Diaz-Mendoza, M., Bideshi, D.K., Federici, B.A., 2012. A 54-kilodalton protein encoded by
pBtoxis is required for parasporal body structural integrity in Bacillus thuringiensis subsp.
israelensis. J. Bacteriol. 194, 15621571.
Donovan, W.P., Donovan, J.C., Slaney, A.C., 2000. Bacillus thuringiensis cryET33 and
cryET34 compositions and uses thereof US Patent 6,063,756.
Du, C., Martin, P.A.W., Nickerson, K.W., 1994. Comparison of disulfide contents and sol-
ubility at alkaline pH of insecticidal and noninsecticidal Bacillus thuringiensis protein crys-
tals. Appl. Environ. Microbiol. 60, 38473853.
Du, C., Nickerson, K.W., 1996. Bacillus thuringiensis HD-73 spores have surface-localized
Cry1Ac toxin: physiological and pathogenic consequences. Appl. Environ. Microbiol.
62, 37223726.
Du, J., Knowles, B.H., Li, J., Ellar, D.J., 1999. Biochemical characterization of Bacillus thur-
ingiensis cytolytic toxins in association with a phospholipid bilayer. Biochem. J.
338, 185193.
Ellis, R.T., Stockhoff, B.A., Stamp, L., Schnepf, H.E., Schwab, G.E., Knuth, M., Russell, J.,
Cardineau, G.A., Narva, K.E., 2002. Novel Bacillus thuringiensis binary insecticidal crystal
proteins active on western corn rootworm, Diabrotica virgifera virgifera LeConte. Appl.
Environ. Microbiol. 68, 11371145.
English, L., Readdy, T.L., 1989. Delta endotoxin inhibits a phosphatase in midgut epithelial
membranes of Heliothis virescens. Insect Biochem. 19, 145152.
Fabrick, J., Oppert, C., Lorenzen, M.D., Morris, K., Oppert, B., Jurat-Fuentes, J.L., 2009.
A novel Tenebrio molitor cadherin is a functional receptor for Bacillus thuringiensis Cry3Aa
toxin. J. Biol. Chem. 284, 1840118410.
Fabrick, J.A., Tabashnik, B.E., 2007. Binding of Bacillus thuringiensis toxin
Cry1Ac to multiple sites of cadherin in pink bollworm. Insect Biochem. Mol. Biol.
37, 97106.
Federici, B.A., Bauer, L.S., 1998. Cyt1Aa protein of Bacillus thuringiensis is toxic to the cot-
tonwood leaf beetle, Chrysomela scripta, and suppresses high levels of resistance to
Cry3Aa. Appl. Environ. Microbiol. 64, 43684371.
Diversity and Mechanisms of Bt Crystal Toxin Action 75
Fernandez-Luna, M.T., Lanz-Mendoza, H., Gill, S.S., Bravo, A., Soberon, M., Miranda-
Rios, J., 2010. An alpha-amylase is a novel receptor for Bacillus thuringiensis ssp. israelensis
Cry4Ba and Cry11Aa toxins in the malaria vector mosquito Anopheles albimanus (Diptera:
Culicidae). Environ. Microbiol. 12, 746757.
Fernandez, L.E., Aimanova, K.G., Gill, S.S., Bravo, A., Soberon, M., 2006. A GPI-anchored
alkaline phosphatase is a functional midgut receptor of Cry11Aa toxin in Aedes aegypti
larvae. Biochem. J. 394, 7784.
Ferre, J., Van Rie, J., 2002. Biochemistry and genetics of insect resistance to Bacillus thur-
ingiensis. Annu. Rev. Entomol. 47, 501533.
Flannagan, R.D., Yu, C.G., Mathis, J.P., Meyer, T.E., Shi, X., Siqueira, H.A., Siegfried, B.-
D., 2005. Identification, cloning and expression of a Cry1Ab cadherin receptor from
European corn borer, Ostrinia nubilalis (Hubner) (Lepidoptera: Crambidae). Insect Bio-
chem. Mol. Biol. 35, 3340.
Flores-Escobar, B., Rodriguez-Magadan, H., Bravo, A., Soberon, M., Gomez, I., 2013. Dif-
ferential role of Manduca sexta aminopeptidase-N and alkaline phosphatase in the mode of
action of Cry1Aa, Cry1Ab, and Cry1Ac toxins from Bacillus thuringiensis. Appl. Environ.
Microbiol. 79, 45434550.
Florez, A.M., Osorio, C., Alzate, O., 2012. Protein engineering of Bacillus thuringiensis
-endotoxins. In: Sansinenea, E. (Ed.), Bacillus thuringiensis Biotechnology. Springer
Netherlands, Netherlands, pp. 93113.
Fortier, M., Vachon, V., Kirouac, M., Schwartz, J.L., Laprade, R., 2005. Differential effects
of ionic strength, divalent cations and pH on the pore-forming activity of Bacillus thur-
ingiensis insecticidal toxins. J. Membr. Biol. 208, 7787.
Fortier, M., Vachon, V., Marceau, L., Schwartz, J.L., Laprade, R., 2007. Kinetics of pore
formation by the Bacillus thuringiensis toxin Cry1Ac. Biochim. Biophys. Acta Biomembr.
1768, 12911298.
Fujii, U., Tanaka, S., Stsuki, M., Hoshino, Y., Morimoto, C., Kotani, T., Harashima, Y.,
Endo, H., Yoshizawa, Y., Sato, R., 2013. Cry1Aa binding to the cadherin receptor does
not require conserved amino acid sequences in the domain II loops. Biosci. Rep. 33 (1),
e00010. http://dx.doi.org/10.01042/BSR20120113.
Gahan, L.J., Gould, F., Heckel, D.G., 2001. Identification of a gene associated with Bt resis-
tance in Heliothis virescens. Science 293, 857860.
Gahan, L.J., Pauchet, Y., Vogel, H., Heckel, D.G., 2010. An ABC transporter mutation is
correlated with insect resistance to Bacillus thuringiensis Cry1Ac toxin. PLoS Genet.
6, e1001248.
Galitsky, N., Cody, V., Wojtczak, A., Ghosh, D., Luft, J.R., Pangborn, W., English, L.,
2001. Structure of the insecticidal bacterial delta-endotoxin Cry3Bb1 of Bacillus thur-
ingiensis. Acta Crystallogr. Sec. D Biol. Crystallogr. 57, 11011109.
Gao, Y., Jurat-Fuentes, J.L., Oppert, B., Fabrick, J.A., Liu, C., Gao, J., Lei, Z., 2011.
Increased toxicity of Bacillus thuringiensis Cry3Aa against Crioceris quatuordecimpunctata,
Phaedon brassicae and Colaphellus bowringi by a Tenebrio molitor cadherin fragment. Pest
Manage. Sci. 67, 10761081.
Garczynski, S.F., Adang, M.J., 1995. Bacillus thuringiensis CryIA(c) d-endotoxin binding ami-
nopeptidase in the Manduca sexta midgut has a glycosyl-phosphatidylinositol anchor.
Insect Biochem. Mol. Biol. 25, 409415.
Garczynski, S.F., Crim, J.W., Adang, M.J., 1991. Identification of putative insect brush bor-
der membrane-binding molecules specific to Bacillus thuringiensis delta-endotoxin by
protein blot analysis. Appl. Environ. Microbiol. 57, 28162820.
Gazit, E., La Rocca, P., Sansom, M.S.P., Shai, Y., 1998. The structure and organization
within the membrane of the helices composing the pore-forming domain of Bacillus thur-
ingiensis d-endotoxin are consistent with an umbrella-like structure of the pore. Proc.
Natl. Acad. Sci. U.S.A. 95, 1228912294.
76 Michael J. Adang et al.
Ge, A.Z., Pfister, R.M., Dean, D.H., 1991. Functional domains of Bacillus thuringiensis insec-
ticidal crystal proteins: refinement of Heliothis virescens and Trichoplusia ni specificity
domains on CryIA(c). J. Biol. Chem. 266, 1795417958.
Gill, M., Ellar, D., 2002. Transgenic Drosophila reveals a functional in vivo receptor for the
Bacillus thuringiensis toxin Cry1Ac1. Insect Mol. Biol. 11, 619625.
Gill, S., Cowles, E.A., Francis, V., 1995. Identification, isolation, and cloning of a Bacillus
thuringiensis CryIAc toxin-binding protein from the midgut of the lepidopteran insect
Heliothis virescens. J. Biol. Chem. 270, 2727727282.
Gill, S.S., Singh, G.J., Hornung, J.M., 1987. Cell membrane interaction of Bacillus thur-
ingiensis subsp. israelensis cytolytic toxins. Infect. Immun. 55, 13001308.
Girard, F., Vachon, V., Prefontaine, G., Marceau, L., Schwartz, J.L., Masson, L.,
Laprade, R., 2009. Helix alpha 4 of the Bacillus thuringiensis Cry1Aa toxin plays a critical
role in the postbinding steps of pore formation. Appl. Environ. Microbiol. 75, 359365.
Goldberg, L.J., Margalit, J., 1977. A bacterial spore demonstrating rapid larvicidal activity
against Anopheles sergentii, Uranotaenia unguicalata, Culex univitattus, Aedes aegyptii and
Culex pipiens. Mosq. News 37, 355358.
Gomez, I., Dean, D.H., Bravo, A., Sober^aon, M., 2003. Molecular basis for Bacillus thur-
ingiensis Cry1Ab toxin specificity: two structural determinants in the Manduca sexta
Bt-R1 receptor interact with loops alpha-8 and 2 in domain II of Cy1Ab toxin.
Biochemistry 42, 1048210489.
Gomez, I., Sanchez, J., Miranda, R., Bravo, A., Soberon, M., 2002. Cadherin-like receptor
binding facilitates proteolytic cleavage of helix -1 in domain I and oligomer pre-pore
formation of Bacillus thuringiensis Cry1Ab toxin. FEBS Lett. 513, 242246.
Gomez, I., Sanchez, J., Miranda, R., Bravo, A., Soberon, M., 2002. Cadherin-like receptor
binding facilitates proteolytic cleavage of helix alpha-1 in domain I and oligomer pre-
pore formation of Bacillus thuringiensis Cry1Ab toxin. FEBS Lett. 513, 242246.
Gomez, I., Sanchez, J., Munoz-Garay, C., Matus, V., Gill, S.S., Soberon, M., Bravo, A.,
2014. Bacillus thuringiensis Cry1A toxins are versatile proteins with multiple modes of
action: two distinct pre-pores are involved in toxicity. Biochem. J. 459, 383396.
Griffitts, J.S., Haslam, S.M., Yang, T., Garczynski, S.F., Mulloy, B., Morris, H., Cremer, P.S.,
Dell, A., Adang, M.J., Aroian, R.V., 2005. Glycolipids as receptors for Bacillus thuringiensis
crystal toxin. Science 307, 922925.
Grochulski, P., Masson, L., Borisova, S., Pusztai-Carey, M., Schwartz, J.L., Brousseau, R.,
Cygler, M., 1995. Bacillus thuringiensis CryIA(a) insecticidal toxin: crystal structure and
channel formation. J. Mol. Biol. 54, 447464.
Groulx, N., Juteau, M., Blunck, R., 2010. Rapid topology probing using fluorescence spec-
troscopy in planar lipid bilayer: the pore-forming mechanism of the toxin Cry1Aa of
Bacillus thuringiensis. J. Gen. Physiol. 136, 497513.
Groulx, N., McGuire, H., Laprade, R., Schwartz, J.L., Blunck, R., 2011. Single molecule
fluorescence study of the Bacillus thuringiensis toxin Cry1Aa reveals tetramerization.
J. Biol. Chem. 286, 4227442282.
Guan, P., Ai, P., Dai, X., Zhang, J., Xu, L., Zhu, J., Li, Q., Deng, Q., Li, S., Wang, S.,
Liu, H., Wang, L., Li, P., Zheng, A., 2012. Complete genome sequence of Bacillus thur-
ingiensis serovar Sichuansis strain MC28. J. Bacteriol. 194, 6975.
Guerchicoff, A., Ugalde, R.A., Rubinstein, C.P., 1997. Identification and characterization
of a previously undescribed cyt gene in Bacillus thuringiensis subsp. israelensis. Appl. Envi-
ron. Microbiol. 63, 27162721.
Guo, S., Li, J., Liu, Y., Song, F., Zhang, J., 2011. The role of DNA binding with the
Cry8Ea1 toxin of Bacillus thuringiensis. FEMS Microbiol. Lett. 317, 203210.
Guo, S., Ye, S., Liu, Y., Wei, L., Xue, J., Wu, H., Song, F., Zhang, J., Wu, X., Huang, D.,
Rao, Z., 2009a. Crystal structure of Bacillus thuringiensis Cry8Ea1: an insecticidal toxin
Diversity and Mechanisms of Bt Crystal Toxin Action 77
Huang, S., Ding, X., Sun, Y., Yang, Q., Xiao, X., Cao, Z., Xia, L., 2012. Proteomic analysis
of Bacillus thuringiensis at different growth phases by using an automated online two-
dimensional liquid chromatography-tandem mass spectrometry strategy. Appl. Environ.
Microbiol. 78, 52705279.
Hughes, A.L., 2014. Evolutionary diversification of aminopeptidase N in Lepidoptera by
conserved clade-specific amino acid residues. Mol. Phylogenet. Evol. 76C, 127133.
Ibargutxi, M.A., Munoz, D., Escudero, I.R.d., Caballero, P., 2008. Interactions between
Cry1Ac, Cry2Ab, and Cry1Fa Bacillus thuringiensis toxins in the cotton pests Helicoverpa
armigera (Hubner) and Earias insulana (Boisduval). Biol. Control 47, 8996.
James, C., 2009. Global Status of Commercialized Biotech/GM Crops, ISAAA Brief.
ISAAA, Ithaca, NY.
Jaquet, F., Hutter, R., Luthy, P., 1987. Specificity of Bacillus thuringiensis -endotoxin. Appl.
Environ. Microbiol. 53, 500504.
Jenkins, J.L., Lee, M.K., Sandagala, S., Adang, M.J., Dean, D.H., 1999. Binding of Bacillus
thuringiensis Cry1Ac toxin to Manduca sexta aminopeptidase-N receptor is not directly
related to toxicity. FEBS Lett. 462, 373376.
Jenkins, J.L., Lee, M.K., Valaitis, A.P., Curtiss, A., Dean, D.H., 2000. Bivalent sequential
binding model of a Bacillus thuringiensis toxin to gypsy moth aminopeptidase
N receptor. J. Biol. Chem. 275, 1442314431.
Jimenez-Juarez, N., Munoz-Garay, C., Gomez, I., Saab-Rincon, G., Damian-Almazo, J.Y.,
Gill, S.S., Soberon, M., Bravo, A., 2007. Bacillus thuringiensis Cry1Ab mutants affecting
oligomer formation are non-toxic to Manduca sexta larvae. J. Biol. Chem.
282, 2122221229.
Jimenez, A.I., Reyes, E.Z., Cancino-Rodezno, A., Bedoya-Perez, L.P., Caballero-Flores,
G.G., Muriel-Millan, L.F., Likitvivatanavong, S., Gill, S.S., Bravo, A., Soberon, M.,
2012. Aedes aegypti alkaline phosphatase ALP1 is a functional receptor of Bacillus thur-
ingiensis Cry4Ba and Cry11Aa toxins. Insect Biochem. Mol. Biol. 42, 683689.
Johnston, P.R., Crickmore, N., 2009. Gut bacteria are not required for the insecticidal activ-
ity of Bacillus thuringiensis toward the tobacco hornworm, Manduca sexta. Appl. Environ.
Microbiol. 75, 50945099.
Jones, G.W., Nielsen-Leroux, C., Yang, Y., Yuan, Z., Dumas, V.F., Monnerat, R.G.,
Berry, C., 2007. A new Cry toxin with a unique two-component dependency from
Bacillus sphaericus. FASEB J. 21, 41124120.
Jurat-Fuentes, J.L., Adang, M.J., 2001. Importance of Cry1 delta-endotoxin domain II loops
for binding specificity in Heliothis virescens (L.). Appl. Environ. Microbiol. 67, 323329.
Jurat-Fuentes, J.L., Adang, M.J., 2004. Characterization of a Cry1Ac-receptor alkaline phos-
phatase in susceptible and resistant Heliothis virescens larvae. Eur. J. Biochem.
271, 31273135.
Jurat-Fuentes, J.L., Adang, M.J., 2006. The Heliothis virescens cadherin protein expressed in
Drosophila S2 cells functions as a receptor for Bacillus thuringiensis Cry1A but not Cry1Fa
toxins. Biochemistry 45, 96889695.
Jurat-Fuentes, J.L., Gahan, L.J., Gould, F.L., Heckel, D.G., Adang, M.J., 2004. The
HevCaLP protein mediates binding specificity of the Cry1A class of Bacillus thuringiensis
toxins in Heliothis virescens. Biochemistry 43, 1429914305.
Jurat-Fuentes, J.L., Jackson, T.A., 2012. Bacterial entomopathogens. In: Vega, F.E.,
Kaya, H.K. (Eds.), Insect Pathology, second ed. Elsevier, San Diego, pp. 265349.
Jurat-Fuentes, J.L., Karumbaiah, L., Jakka, S.R., Ning, C., Liu, C., Wu, K., Jackson, J.,
Gould, F., Blanco, C., Portilla, M., Perera, O., Adang, M., 2011. Reduced levels of
membrane-bound alkaline phosphatase are common to lepidopteran strains resistant
to Cry toxins from Bacillus thuringiensis. PLoS One 6, e17606.
Karlova, R., Weemen-Hendriks, M., Naimov, S., Ceron, J., Dukiandjiev, S., de Maagd,
R.A., 2005. Bacillus thuringiensis delta-endotoxin Cry1Ac domain III enhances activity
Diversity and Mechanisms of Bt Crystal Toxin Action 79
against Heliothis virescens in some, but not all Cry1-Cry1Ac hybrids. J. Invertebr. Pathol.
88, 169172.
Katayama, H., Kusaka, Y., Yokota, H., Akao, T., Kojima, M., Nakamura, O., Mekada, E.,
Mizuki, E., 2007. Parasporin-1, a novel cytotoxic protein from Bacillus thuringiensis,
induces Ca2 + influx and a sustained elevation of the cytoplasmic Ca2 + concentration
in toxin-sensitive cells. J. Biol. Chem. 282, 77427752.
Kauer, R., Sharma, A., Gupta, D., Kalita, M., Bhatnagar, R.K., 2014. Bacillus thuringiensis
toxin, Cry1C interacts with 128HLHFHLP134 region of aminopeptidase N of agricul-
tural pest, Spodoptera litura. Process Biochem. 49, 688696.
Keeton, T.P., Bulla Jr., L.A., 1997. Ligand specificity and affinity of BT-R1, the Bacillus thur-
ingiensis Cry1A toxin receptor from Manduca sexta, expressed in mammalian and insect
cell cultures. Appl. Environ. Microbiol. 63, 34193425.
Kim, H.S., Saitoh, H., Yamashita, S., Akao, T., Park, Y.S., Maeda, M., Tanaka, R.,
Mizuki, E., Ohba, M., 2003. Cloning and characterization of two novel crystal protein
genes from a Bacillus thuringiensis serovar dakota strain. Curr. Microbiol. 46, 3338.
Kitada, S., Abe, Y., Maeda, T., Shimada, H., 2009. Parasporin-2 requires GPI-anchored
proteins for the efficient cytocidal action to human hepatoma cells. Toxicology
264, 8088.
Knight, P.J., Knowles, B.H., Ellar, D.J., 1995. Molecular cloning of an insect aminopeptidase
N that serves as a receptor for Bacillus thuringiensis CryIA(c) toxin. J. Biol. Chem.
270, 1776517770.
Knight, P.J.K., Knowles, B.H., Ellar, D.J., 1994. The receptor for Bacillus thuringiensis
CryIA(c) delta-endotoxin in the brush border membrane is aminopepidase N. Mol.
Microbiol. 11, 429436.
Knowles, B.H., 1994. Mechanism of action of Bacillus thuringiensis insecticidal -endotoxins.
Adv. Insect Physiol. 24, 275308.
Knowles, B.H., Blatt, M.R., Tester, M., Horsnell, J.M., Carroll, J., Menestrina, G., Ellar, D.J.,
1989. A cytolytic delta-endotoxin from Bacillus thuringiensis var. israelensis forms cation-
selective channels in planar lipid bilayers. FEBS Lett. 244, 259262.
Koni, P.A., Ellar, D.J., 1994. Biochemical characterization of Bacillus thuringiensis cytolytic
delta-endotoxins. Microbiology 140, 18691880.
Kramer, K.J., Muthukrishnan, S., 1997. Insect chitinases: molecular biology and potential use
as biopesticides. Insect Biochem. Mol. Biol. 27, 887900.
Krishnamoorthy, M., Jurat-Fuentes, J.L., McNall, R.J., Andacht, T., Adang, M.J., 2007.
Identification of novel Cry1Ac binding proteins in midgut membranes from Heliothis
virescens using proteomic analyses. Insect Biochem. Mol. Biol. 37, 189201.
Krishnan, V., 2013. Investigation of Parasporins, the Cytotoxic Proteins from the Bacterium
Bacillus thuringiensis. University of Sussex, United Kingdom.
Kumar, A.S., Aronson, A.I., 1999. Analysis of mutations in the pore-forming region
essential for insecticidal activity of a Bacillus thuringiensis delta-endotoxin. J. Bacteriol.
181, 61036107.
Lailak, C., Khaokhiew, T., Promptmas, C., Promdonkoy, B., Pootanakit, K.,
Angsuthanasombat, C., 2013. Bacillus thuringiensis Cry4Ba toxin employs two
receptor-binding loops for synergistic interactions with Cyt2Aa2. Biochem. Biophys.
Res. Commun. 435, 216221.
Lee, K.Y., Kang, E.Y., Park, S., Ahn, S.K., Yoo, K.H., Kim, J.Y., Lee, H.H., 2006. Mass
spectrometric sequencing of endotoxin proteins of Bacillus thuringiensis ssp. konkukian
extracted from polyacrylamide gels. Proteomics 6, 15121517.
Lee, M.K., Jenkins, J.L., You, T.H., Curtiss, A., Son, J.J., Adang, M.J., Dean, D.H., 2001.
Mutations at the arginine residues in alpha8 loop of Bacillus thuringiensis delta-endotoxin
Cry1Ac affect toxicity and binding to Manduca sexta and Lymantria dispar aminopeptidase
N. FEBS Lett. 497, 108112.
80 Michael J. Adang et al.
Lee, M.K., Rajamohan, F., Jenkins, J.L., Curtiss, A.S., Dean, D.H., 2000. Role of two
arginine residues in domain II, loop 2 of Cry1Ab and Cry1Ac Bacillus thuringiensis
delta-endotoxin in toxicity and binding to Manduca sexta and Lymantria dispar aminopep-
tidase N. Mol. Microbiol. 38, 289298.
Lee, M.K., Young, B.A., Dean, D.H., 1995. Domain III exchanges of Bacillus thuringiensis
Cry1A toxins affect binding to different gypsy moth midgut receptors. Biochem.
Biophys. Res. Commun. 216, 306312.
Lee, X., Johnston, R.A., Rose, D.R., Young, N.M., 1989. Crystallization and preliminary
X-ray diffraction studies of the complex of Maclura pomifera agglutinin with the disaccha-
ride Gal beta 1-3GalNAc. J. Mol. Biol. 210, 685686.
Li, H., Olson, M., Lin, G., Hey, T., Tan, S.Y., Narva, K.E., 2013. Bacillus thuringiensis
Cry34Ab1/Cry35Ab1 interactions with western corn rootworm midgut membrane
binding sites. PLoS One 8, e53079.
Li, J., Derbyshire, D.J., Promdonkoy, B., Ellar, D.J., 2001. Structural implications for the
transformation of the Bacillus thuringiensis delta-endotoxins from water-soluble to
membrane-inserted forms. Biochem. Soc. Trans. 29, 571577.
Li, J., Koni, P.A., Ellar, D.J., 1996. Structure of the mosquitocidal delta-endotoxin CytB
from Bacillus thuringiensis sp kyushuensis and implications for membrane pore formation.
J. Mol. Biol. 257, 129152.
Li, J.D., Carroll, J., Ellar, D.J., 1991. Crystal structure of insecticidal delta-endotoxin from
Bacillus thuringiensis at 2.5 A resolution. Nature 353 (6347), 815821.
Liang, Y., Patel, S.S., Dean, D.H., 1995. Irreversible binding kinetics of Bacillus thuringiensis
CryIA delta-endotoxins to gypsy moth brush border membrane vesicles is directly cor-
related to toxicity. J. Biol. Chem. 270, 2471924724.
Likitvivatanavong, S., Chen, J., Bravo, A., Soberon, M., Gill, S.S., 2011. Cadherin, alkaline
phosphatase, and aminopeptidase N as receptors of Cry11Ba toxin from Bacillus
thuringiensis subsp. jegathesan in Aedes aegypti. Appl. Environ. Microbiol. 77, 2431.
Lopez-Diaz, J.A., Canton, P.E., Gill, S.S., Soberon, M., Bravo, A., 2013. Oligomerization is
a key step in Cyt1Aa membrane insertion and toxicity but not necessary to synergize
Cry11Aa toxicity in Aedes aegypti larvae. Environ. Microbiol. 15, 30303039.
Lu, H., Rajamohan, F., Dean, D.H., 1994. Identification of amino acid residues of Bacillus
thuringiensis -endotoxin CryIAa associated with membrane binding and toxicity to
Bombyx mori. J. Bacteriol. 176, 55545559.
Luo, K., Sangadala, S., Masson, L., Mazza, A., Brousseau, R., Adang, M.J., 1997. The
Heliothis virescens 170-kDa aminopeptidase functions as Receptor A by mediating spe-
cific Bacillus thuringiensis Cry1A -endotoxin binding and pore formation. Insect Bio-
chem. Mol. Biol. 27, 735743.
Ma, G., Rahman, M.M., Grant, W., Schmidt, O., Asgari, S., 2012. Insect tolerance to the
crystal toxins Cry1Ac and Cry2Ab is mediated by the binding of monomeric toxin to
lipophorin glycolipids causing oligomerization and sequestration reactions. Dev. Comp.
Immunol. 37, 184192.
Manasherob, R., Itsko, M., Sela-Baranes, N., Ben-Dov, E., Berry, C., Cohen, S.,
Zaritsky, A., 2006. Cyt1Ca from Bacillus thuringiensis subsp. israelensis: production
in Escherichia coli and comparison of its biological activities with those of other Cyt-like
proteins. Microbiology 152, 26512659.
Manceva, S.D., Pusztai-Carey, M., Russo, P.S., Butko, P., 2005. A detergent-like mecha-
nism of action of the cytolytic toxin Cyt1A from Bacillus thuringiensis var. israelensis.
Biochemistry 44, 589597.
Martins, E.S., Monnerat, R.G., Queiroz, P.R., Dumas, V.F., Braz, S.V., de Souza Aguiar,
R.W., Gomes, A.C., Sanchez, J., Bravo, A., Ribeiro, B.M., 2010. Midgut GPI-
anchored proteins with alkaline phosphatase activity from the cotton boll weevil
Diversity and Mechanisms of Bt Crystal Toxin Action 81
(Anthonomus grandis) are putative receptors for the Cry1B protein of Bacillus thuringiensis.
Insect Biochem. Mol. Biol. 40, 138145.
Masson, L., Lu, Y.J., Mazza, A., Brousseau, R., Adang, M.J., 1995. The CryIA(c) receptor
purified from Manduca sexta displays multiple specificities. J. Biol. Chem.
270, 2030920315.
Masson, L., Mazza, A., Sangadala, S., Adang, M.J., Brousseau, R., 2002a. Polydispersity of
Bacillus thuringiensis Cry1 toxins in solution and its effect on receptor binding kinetics.
Biochim. Biophys. Acta 1594, 266275.
Masson, L., Schwab, G., Mazza, A., Brousseau, R., Potvin, L., Schwartz, J.L., 2004. A novel
Bacillus thuringiensis (PS149B1) containing a Cry34Ab1/Cry35Ab1 binary toxin specific
for the western corn rootworm Diabrotica virgifera virgifera LeConte forms ion channels in
lipid membranes. Biochemistry 43, 1234912357.
Masson, L., Tabashnik, B.E., Liu, Y.B., Schwartz, J.L., 1999. Helix 4 of the Bacillus thur-
ingiensis Cry1Aa toxin lines the lumen of the ion channel. J. Biol. Chem.
274, 3199632000.
Masson, L., Tabashnik, B.E., Mazza, A., Pr^aefontaine, G., Potvin, L., Brousseau, R.,
Schwartz, J.L., 2002b. Mutagenic analysis of a conserved region of domain III in the
Cry1Ac toxin of Bacillus thuringiensis. Appl. Environ. Microbiol. 68, 194200.
Matsushima-Hibiya, Y., Watanabe, M., Hidari, K.I., Miyamoto, D., Suzuki, Y., Kasama, T.,
Koyama, K., Sugimura, T., Wakabayashi, K., 2003. Identification of glycosphingolipid
receptors for pierisin-1, a guanine-specific ADP-ribosylating toxin from the cabbage
butterfly. J. Biol. Chem. 278, 99729978.
McNall, R.J., Adang, M.J., 2003. Identification of novel Bacillus thuringiensis Cry1Ac binding
proteins in Manduca sexta midgut through proteomic analysis. Insect Biochem. Mol. Biol.
33, 9991010.
Mizuki, E., Ohba, M., Akao, T., Yamashita, S., Saitoh, H., Park, Y.S., 1999. Unique activity
associated with non-insecticidal Bacillus thuringiensis parasporal inclusions: in vitro cell-
killing action on human cancer cells. J. Appl. Microbiol. 86, 477486.
Morin, S., Biggs, R.W., Sisteron, M.S., Shriver, L., Ellers-Kirk, C., Higginson, D.,
Holley, D., Gahan, L.J., Heckel, D.G., Carriere, Y., Dennehy, T.J., Brown, J.K.,
Tabashnik, B.E., 2003. Three cadherin alleles associated with resistance to Bacillus thur-
ingiensis in pink bollworm. Proc. Natl. Acad. Sci. U.S.A. 100, 50045009.
Morse, R.J., Yamamoto, T., Stroud, R.M., 2001. Structure of Cry2Aa suggests an unex-
pected receptor binding epitope. Structure (Camb.) 9, 409417.
Nagamatsu, Y., Koike, T., Sasaki, K., Yoshimoto, A., Furukawa, Y., 1999. The cadherin-
like protein is essential to specificity determination and cytotoxic action of the Bacillus
thuringiensis insecticidal CryIAa toxin. FEBS Lett. 460, 385390.
Nagamatsu, Y., Okamura, S., Saitou, H., Akao, T., Mizuki, E., 2010. Three Cry toxins in
two types from Bacillus thuringiensis strain M019 preferentially kill human hepatocyte can-
cer and uterus cervix cancer cells. Biosci. Biotechnol. Biochem. 74, 494498.
Naimov, S., Boncheva, R., Karlova, R., Dukiandjiev, S., Minkov, I., de Maagd, R.A., 2008.
Solubilization, activation, and insecticidal activity of Bacillus thuringiensis serovar tho-
mpsoni HD542 crystal proteins. Appl. Environ. Microbiol. 74, 71457151.
Naimov, S., Valkova, R., Dukiandjiev, S., Minkov, I., de Maagd, R.A., 2011. Carboxy-
terminal extension effects on crystal formation and insecticidal properties of Cry15Aa.
J. Invertebr. Pathol. 108, 5658.
Nair, M.S., Dean, D.H., 2008. All domains of Cry1A toxins insert into insect brush border
membranes. J. Biol. Chem. 283, 2632426331.
Ning, C., Wu, K., Liu, C., Gao, Y., Jurat-Fuentes, J.L., Gao, X., 2010. Characterization of a
Cry1Ac toxin-binding alkaline phosphatase in the midgut from Helicoverpa armigera
(Hubner) larvae. J. Insect Physiol. 56, 666672.
82 Michael J. Adang et al.
Obata, F., Kitami, M., Inoue, Y., Atsumi, S., Yoshizawa, Y., Sato, R., 2009. Analysis of the
region for receptor binding and triggering of oligomerization on Bacillus thuringiensis
Cry1Aa toxin. FEBS J. 276, 59495959.
Ochoa-Campuzano, C., Martnez-Ramrez, A.C., Contreras, E., Rausell, C., Real, M.D.,
2013. Prohibitin, an essential protein for Colorado potato beetle larval viability, is rele-
vant to Bacillus thuringiensis Cry3Aa toxicity. Pestic. Biochem. Physiol. 107, 299308.
Ochoa-Campuzano, C., Real, M.D., Martnez-Ramrez, A.C., Bravo, A., Rausell, C.,
2007. An ADAM metalloprotease is a Cry3Aa Bacillus thuringiensis toxin receptor. Bio-
chem. Biophys. Res. Commun. 362, 437442.
Okumura, S., Saitoh, H., Ishikawa, T., Inouye, K., Mizuki, E., 2011. Mode of action of
parasporin-4, a cytocidal protein from Bacillus thuringiensis. Biochim. Biophys. Acta
1808, 14761482.
Oppert, B., 1999. Protease interactions with Bacillus thuringiensis insecticidal toxins. Arch.
Insect Biochem. Physiol. 42, 112.
Ounjai, P., Unger, V.M., Sigworth, F.J., Angsuthanasombat, C., 2007. Two conformational
states of the membrane-associated Bacillus thuringiensis Cry4Ba -endotoxin complex
revealed by electron crystallography: implications for toxin-pore formation. Biochem.
Biophys. Res. Commun. 361, 890895.
Pacheco, S., Gomez, I., Arenas, I., Saab-Rincon, G., Rodrguez-Almazan, C., Gill, S.S.,
Bravo, A., Soberon, M., 2009a. Domain II loop 3 of Bacillus thuringiensis Cry1Ab toxin
is involved in a ping pong binding mechanism with Manduca sexta aminopeptidase-N
and cadherin receptors. J. Biol. Chem. 284, 3275032757.
Pacheco, S., Gomez, I., Gill, S.S., Bravo, A., Soberon, M., 2009b. Enhancement of insec-
ticidal activity of Bacillus thuringiensis Cry1A toxins by fragments of a toxin-binding
cadherin correlates with oligomer formation. Peptides 30, 583588.
Pandian, G.N., Ishikawa, T., Togashi, M., Shitomi, Y., Haginoya, K., Yamamoto, S.,
Nishiumi, T., Hori, H., 2008. Bombyx mori midgut membrane protein P252, which
binds to Bacillus thuringiensis Cry1A, is a chlorophyllide-binding protein, and the resulting
complex has antimicrobial activity. Appl. Environ. Microbiol. 74, 13241331.
Pandian, G.N., Ishikawa, T., Vaijayanthi, T., Hossain, D.M., Yamamoto, S., Nishiumi, T.,
Angsuthanasombat, C., Haginoya, K., Mitsui, T., Hori, H., 2010. Formation of macro-
molecule complex with Bacillus thuringiensis Cry1A toxins and chlorophyllide binding
252-kDa lipocalin-like protein locating on Bombyx mori midgut membrane.
J. Membr. Biol. 237, 125136.
Pardo-Lopez, L., Gomez, I., Rausell, C., Sanchez, J., Soberon, M., Bravo, A., 2006. Struc-
tural changes of the Cry1Ac oligomeric pre-pore from Bacillus thuringiensis induced by
N-acetylgalactosamine facilitates toxin membrane insertion. Biochemistry
45, 1032910336.
Pardo-Lopez, L., Soberon, M., Bravo, A., 2013. Bacillus thuringiensis insecticidal three-
domain Cry toxins: mode of action, insect resistance and consequences for crop protec-
tion. FEMS Microbiol. Rev. 37, 322.
Pardo-Lopez, L., Soberon, M., Bravo, A., 2012. Bacillus thuringiensis insecticidal three-
domain Cry toxins: mode of action, insect resistance and consequences for crop protec-
tion. FEMS Microbiol. Rev. 37, 322.
Park, Y., Kim, Y., 2013. RNA interference of cadherin gene expression in Spodoptera
exigua reveals its significance as a specific Bt target. J. Invertebr. Pathol. 114, 285291.
Peng, D., Xu, X., Ye, W., Yu, Z., Sun, M., 2010. Helicoverpa armigera cadherin fragment
enhances Cry1Ac insecticidal activity by facilitating toxin-oligomer formation. Appl.
Microbiol. Biotechnol. 85, 10331040.
Perez, C., Munoz-Garay, C., Portugal, L.C., Sanchez, J., Gill, S.S., Soberon, S., Bravo, A.,
2007. Bacillus thuringiensis ssp. israelensis Cyt1Aa enhances activity of Cry11Aa toxin by
Diversity and Mechanisms of Bt Crystal Toxin Action 83
Rodriguez-Almazan, C., Ruiz de Escudero, I., Canton, P.E., Munoz-Garay, C., Perez, C.,
Gill, S.S., Soberon, M., Bravo, A., 2011. The amino- and carboxyl-terminal fragments of
the Bacillus thuringiensis Cyt1Aa toxin have differential roles in toxin oligomerization and
pore formation. Biochemistry 50, 388396.
Rosa, J.C., De Oliveira, P.S., Garratt, R., Beltramini, L., Resing, K., Roque-Barreira, M.C.,
Greene, L.J., 1999. KM+, a mannose-binding lectin from Artocarpus integrifolia: amino
acid sequence, predicted tertiary structure, carbohydrate recognition, and analysis of
the beta-prism fold. Protein Sci. 8, 1324.
Saengwiman, S., Aroonkesorn, A., Dedvisitsakul, P., Sakdee, S., Leetachewa, S.,
Angsuthanasombat, C., Pootanakit, K., 2011. In vivo identification of Bacillus
thuringiensis Cry4Ba toxin receptors by RNA interference knockdown of
glycosylphosphatidylinositol-linked aminopeptidase N transcripts in Aedes aegypti larvae.
Biochem. Biophys. Res. Commun. 407, 708713.
Sakurai, J., Nagahama, M., Oda, M., Tsuge, H., Kobayashi, K., 2009. Clostridium perfringens
iota-toxin: structure and function. Toxins 1, 208228.
Sangadala, S., Walters, F.S., English, L.H., Adang, M.J., 1994. A mixture of Manduca sexta
aminopeptidase and phosphatase enhances Bacillus thuringiensis insecticidal CryIA(c)
toxin binding and 86Rb(+)-K+ efflux in vitro. J. Biol. Chem. 269, 1008810092.
Sankaranarayanan, R., Sekar, K., Banerjee, R., Sharma, V., Surolia, A., Vijayan, M., 1996.
A novel mode of carbohydrate recognition in jacalin, a Moraceae plant lectin with a beta-
prism fold. Nat. Struct. Biol. 3, 596603.
Sarkar, A., Hess, D., Mondal, H.A., Banerjee, S., Sharma, H.C., Das, S., 2009.
Homodimeric alkaline phosphatase located at Helicoverpa armigera midgut, a putative
receptor of Cry1Ac contains alpha-GalNAc in terminal glycan structure as interactive
epitope. J. Proteome Res. 8, 18381848.
Schnepf, E., Crickmore, N., Van Rie, J., Lereclus, D., Baum, J., Feitelson, J., Zeigler, D.R.,
Dean, D.H., 1998. Bacillus thuringiensis and its pesticidal crystal proteins. Microbiol.
Mol. Biol. Rev. 62, 775806.
Schnepf, H.E., Lee, S., Dojillo, J., Burmeister, P., Fencil, K., Morera, L., Nygaard, L.,
Narva, K.E., Wolt, J.D., 2005. Characterization of Cry34/Cry35 binary insecticidal pro-
teins from diverse Bacillus thuringiensis strain collections. Appl. Environ. Microbiol.
71, 17651774.
Schwartz, J.L., Juteau, M., Grochulski, P., Cygler, M., Pr^aefontaine, G., Brousseau, R.,
Masson, L., 1997a. Restriction of intramolecular movements within the Cry1Aa toxin
molecule of Bacillus thuringiensis through disulfide bond engineering. FEBS Lett.
410, 397402.
Schwartz, J.L., Lu, Y.J., Soehnlein, P., Brousseau, R., Masson, L., Laprade, R., Adang, M.J.,
1997b. Ion channels formed in planar lipid bilayers by Bacillus thuringiensis toxins in the
presence of Manduca sexta midgut receptors. FEBS Lett. 412, 270276.
Schwartz, J.L., Potvin, L., Chen, X.J., Brousseau, R., Laprade, R., Dean, D.H., 1997c.
Single-site mutations in the conserved alternating-arginine region affect ionic channels
formed by CryIAa, a Bacillus thuringiensis toxin. Appl. Environ. Microbiol.
63, 39783984.
Sengupta, A., Sarkar, A., Priya, P., Ghosh Dastidar, S., Das, S., 2013. New insight to
structure-function relationship of GalNAc mediated primary interaction between insec-
ticidal Cry1Ac toxin and HaALP receptor of Helicoverpa armigera. PLoS One
8, e78249.
Sharma, A., Chandran, D., Singh, D.D., Vijayan, M., 2007. Multiplicity of carbohydrate-
binding sites in beta-prism fold lectins: occurrence and possible evolutionary implica-
tions. J. Biosci. 32, 10891110.
Sher, D., Fishman, Y., Zhang, M., Lebendiker, M., Gaathon, A., Mancheno, J.M.,
Zlotkin, E., 2005. Hydralysins, a new category of beta-pore-forming toxins in cnidaria.
J. Biol. Chem. 280, 2284722855.
Diversity and Mechanisms of Bt Crystal Toxin Action 85
Sivakumar, S., Rajagopal, R., Venkatesh, G.R., Srivastava, A., Bhatnagar, R.K., 2007.
Knockdown of aminopeptidase-N from Helicoverpa armigera larvae and in transfected
Sf21 cells by RNA interference reveals its functional interaction with Bacillus thuringiensis
insecticidal protein Cry1Ac. J. Biol. Chem. 282, 73127319.
Smedley, D.P., Ellar, D.J., 1996. Mutagenesis of three surface-exposed loops of a Bacillus thur-
ingiensis insecticidal toxin reveals residues important for toxicity, receptor recognition
and possibly membrane insertion. Microbiology 142, 16171624.
Soberon, M., Lopez-Diaz, J.A., Bravo, A., 2013a. Cyt toxins produced by Bacillus thur-
ingiensis: a protein fold conserved in several pathogenic microorganisms. Peptides
41, 8793.
Soberon, M., Pardo-Lopez, L., Lopez, I., Gomez, I., Tabashnik, B.E., Bravo, A., 2007. Engi-
neering modified Bt toxins to counter insect resistance. Science 318, 16401642.
Staples, N., Ellar, D., Crickmore, N., 2001. Cellular localization and characterization of the
Bacillus thuringiensis Orf2 crystallization factor. Curr. Microbiol. 42, 388392.
Tabashnik, B.E., Huang, F., Ghimire, M.N., Leonard, B.R., Siegfried, B.D.,
Rangasamy, M., Yang, Y., Wu, Y., Gahan, L.J., Heckel, D.G., Bravo, A.,
Soberon, M., 2011. Efficacy of genetically modified Bt toxins against insects with differ-
ent genetic mechanisms of resistance. Nat. Biotechnol. 29, 11281131.
Tanaka, S., Miyamoto, K., Noda, H., Jurat-Fuentes, J.L., Yoshizawa, Y., Endo, H., Sato, R.,
2013. The ATP-binding cassette transporter subfamily C member 2 in Bombyx mori larvae
is a functional receptor for Cry toxins from Bacillus thuringiensis. FEBS J. 280, 17821794.
Terra, W.R., Ferreira, C., 1994. Insect digestive enzymes: properties, compartmentalization
and function. Comp. Biochem. Physiol. 109B, 162.
Terra, W.R., Ferreira, C., Jordao, B.P., Dillon, R.J., 1996. Digestive enzymes.
In: Lehane, M.J., Billingsley, P.F. (Eds.), Biology of the Insect Midgut. Chapman &
Hall, London, pp. 153194.
Tetreau, G., Bayyareddy, K., Jones, C.M., Stalinski, R., Riaz, M.A., Paris, M., David, J.P.,
Adang, M.J., Despres, L., 2012. Larval midgut modifications associated with Bti resis-
tance in the yellow fever mosquito using proteomic and transcriptomic approaches.
BMC Genomics 13, 248.
Thiery, I., Hamon, S., 1998. Bacterial control of mosquito larvae: investigation of stability of
Bacillus thuringiensis var. israelensis and Bacillus sphaericus standard powders. J. Am. Mosq.
Control Assoc. 14, 472476.
Thomas, W.E., Ellar, D.J., 1983a. Bacillus thuringiensis var israelensis crystal delta-endotoxin:
effects on insect and mammalian cells in vitro and in vivo. J. Cell Sci. 60, 181197.
Thomas, W.E., Ellar, D.J., 1983b. Mechanism of action of Bacillus thuringiensis var israelensis
insecticidal delta-endotoxin. FEBS Lett. 154, 362368.
Tiewsiri, K., Wang, P., 2011. Differential alteration of two aminopeptidases N associated
with resistance to Bacillus thuringiensis toxin Cry1Ac in cabbage looper. Proc. Natl. Acad.
Sci. U.S.A. 108, 1403714042.
Tigue, N.J., Jacoby, J., Ellar, D.J., 2001. The alpha-helix 4 residue, Asn135, is involved in the
oligomerization of Cry1Ac1 and Cry1Ab5 Bacillus thuringiensis toxins. Appl. Environ.
Microbiol. 67, 57155720.
Torres, J., Lin, X., Boonserm, P., 2008. A trimeric building block model for Cry toxins
in vitro ion channel formation. Biochim. Biophys. Acta Biomembr. 1778, 392397.
Tyrell, D.J., Bulla Jr., L.A., Andrews Jr., R.E., Kramer, K.J., Davidson, L.I., Nordin, P.,
1981. Comparative biochemistry of entomocidal parasporal crystals of selected Bacillus
thuringiensis strains. J. Bacteriol. 145, 10521062.
Vachon, V., Laprade, R., Schwartz, J.L., 2012. Current models of the mode of action of Bacil-
lus thuringiensis insecticidal crystal proteins: a critical review. J. Invertebr. Pathol.
111, 112.
Vadlamudi, R.K., Weber, E., Ji, I., Ji, T.H., Bulla Jr., L.A., 1995. Cloning and expression of a
receptor for an insecticidal toxin of Bacillus thuringiensis. J. Biol. Chem. 270, 54905494.
86 Michael J. Adang et al.
Yamaguchi, T., Sahara, K., Bando, H., Asano, S.-I., 2010. Intramolecular proteolytic
nicking of Bacillus thuringiensis Cry8Da toxin in BBMVs of Japanese beetle.
J. Invertebr. Pathol. 105, 243247.
Yamamoto, T., Iizuka, T., Aronson, J.N., 1983. Mosquitocidal protein of Bacillus thur-
ingiensis subsp. israelensis: identification and partial isolation of the protein. Curr.
Microbiol. 9, 279284.
Yamashita, S., Katayama, H., Saitoh, H., Akao, T., Park, Y.S., Mizuki, E., Ohba, M., Ito, A.,
2005. Typical three-domain Cry proteins of Bacillus thuringiensis strain A1462 exhibit
cytocidal activity on limited human cancer cells. J. Biochem. 138, 663672.
Yang, Y., Zhu, Y.C., Ottea, J., Husseneder, C., Leonard, B.R., Abel, C., Huang, F., 2010.
Molecular characterization and RNA interference of three midgut aminopeptidase
N isozymes from Bacillus thuringiensis-susceptible and -resistant strains of sugarcane borer,
Diatraea saccharalis. Insect Biochem. Mol. Biol. 40, 592603.
Ye, W., Zhu, L., Liu, Y., Crickmore, N., Peng, D., Ruan, L., Sun, M., 2012. Mining new
crystal protein genes from Bacillus thuringiensis on the basis of mixed plasmid-enriched
genome sequencing and a computational pipeline. Appl. Environ. Microbiol.
78, 47954801.
Zhang, B., Liu, M., Yang, Y., Yuan, Z., 2006. Cytolytic toxin Cyt1Aa of Bacillus thuringiensis
synergizes the mosquitocidal toxin Mtx1 of Bacillus sphaericus. Biosci. Biotechnol. Bio-
chem. 70, 21992204.
Zhang, H., Wu, S., Yang, Y., Tabashnik, B.E., Wu, Y., 2012. Non-recessive Bt toxin resis-
tance conferred by an intracellular cadherin mutation in field-selected populations of cot-
ton bollworm. PLoS One 7, e53418.
Zhang, R., Hua, G., Andacht, T.M., Adang, M.J., 2008. A 106-kDa aminopeptidase is a
putative receptor for Bacillus thuringiensis Cry11Ba toxin in the mosquito Anopheles
gambiae. Biochemistry 47, 1126311272.
Zhang, R., Hua, G., Urbauer, J.L., Adang, M.J., 2010. Synergistic and inhibitory effects of
aminopeptidase peptides on Bacillus thuringiensis Cry11Ba toxicity in the mosquito
Anopheles gambiae. Biochemistry 49, 85128519.
Zhang, S., Cheng, H., Gao, Y., Wang, G., Liang, G., Wu, K., 2009. Mutation of an ami-
nopeptidase N gene is associated with Helicoverpa armigera resistance to Bacillus thur-
ingiensis Cry1Ac toxin. Insect Biochem. Mol. Biol. 39, 421429.
Zhang, X., Candas, M., Griko, N.B., Rose-Young, L., Bulla, L.A., 2005. Cytotoxicity of
Bacillus thuringiensis Cry1Ab toxin depends on specific binding of the toxin to the
cadherin receptor BT-R(1) expressed in insect cells. Cell Death Differ. 12, 14071416.
Zhou, X., Konkel, M.E., Call, D.R., 2009. Type III secretion system 1 of Vibrio para-
haemolyticus induces oncosis in both epithelial and monocytic cell lines. Microbiology
155, 837851.
Zhuang, M., Oltean, D.I., Gomez, I., Pullikuth, A.K., Soberon, M., Bravo, A., Gill, S.S.,
2002. Heliothis virescens and Manduca sexta lipid rafts are involved in Cry1A toxin binding
to the midgut epithelium and subsequent pore formation. J. Biol. Chem.
277, 1386313872.
Zuniga-Navarrete, F., Gomez, I., Pena, G., Bravo, A., Soberon, M., 2013. A Tenebrio molitor
GPI-anchored alkaline phosphatase is involved in binding of Bacillus thuringiensis Cry3Aa
to brush border membrane vesicles. Peptides 41, 8186.
CHAPTER THREE
Contents
1. Introduction 90
1.1 Background 90
1.2 General features and strains 91
1.3 The relevance of L. sphaericus as a mosquito-control agent 93
2. Toxins and Mode of Action 95
2.1 Spectrum of action 95
2.2 Binary toxin 97
2.3 Cry48/Cry49 104
2.4 Mosquitocidal toxin 1 104
2.5 Other Mtx toxins 107
2.6 Sphaericolysin 108
2.7 S-layer proteins 109
2.8 Safety issues 109
3. Receptors of the Binary Toxin 112
3.1 Binding of the binary toxin to larvae midgut 112
3.2 Receptors 114
3.3 Comparative analysis of the Cqm1 and Aam1 -glucosidases 119
4. Applications for Mosquito Control 121
4.1 Field trials 121
4.2 Factors affecting field performance 125
4.3 Trials against the vectors of lymphatic filariasis 126
4.4 Recent large-scale trials 127
4.5 Operational use in mosquito-control programmes 129
5. Resistance 130
5.1 Factors involved in the selection of resistance 130
5.2 Laboratory and field reports 131
5.3 Mechanisms and inheritance of resistance 135
Abstract
Lysinibacillus sphaericus (Ls) strains that produce insecticidal proteins show high activity
against mosquito larvae. The most active of these is the binary (Bin) toxin that acts fol-
lowing ingestion and, after midgut processing and binding to specific receptors, pro-
vokes cytopathological effects and leads to larval death. Bin toxin displays specific action
against some species of medical importance (e.g. Culex and Anopheles) and it is safe to
non-target organisms. These features have led to the production of biolarvicides based
on this bacterium and its effectiveness to control mosquito larvae has been widely
related in the literature. The field utilisation of Ls has also shown that resistance could
be selected among exposed populations and the mechanisms and genes involved in
this process have been described. Management strategies can be successfully
employed to avoid resistance and Ls can be used within integrated programmes as
a selective and efficient agent to control mosquitoes.
1. INTRODUCTION
1.1. Background
The utilisation of entomopathogenic bacteria for insect control started in the
1960s with the discovery and development of Bacillus thuringiensis (Bt) vari-
eties that produced insecticidal proteins active against agricultural insect
pests. The B. thuringiensis serovar. israelensis (Bti) discovered by Goldberg
and Margalit (1978) was the first serotype identified as active against Diptera
larvae (de Barjac, 1978). This entomopathogenic bacterium enjoyed a rapid
development from the characterisation of its properties to field utilisation
(Becker, 1997; Guillet et al., 1990; Margalit and Dean, 1985), mainly
because of the serious resistance problems encountered by synthetic insec-
ticides in vector-control programmes during that period. The second
mosquitocidal bacterium Lysinibacillus sphaericus (Ls), previously designated
as B. sphaericus, was identified by Neide in 1904 (Neide, 1904). Character-
isation of this species as a mosquito pathogen was initiated by Kellen, much
Microbial Toxins for Mosquito Control 91
later, when a toxic strain was isolated from cadavers of Culiseta incidens larvae
(Kellen et al., 1965). The Kellen (K) strain displayed a low level of toxicity
and did not attract much interest for its development as a control agent. The
discovery by Singer (1973), of the SSII-1 strain, that displayed a higher activ-
ity than the K strain renewed the interest in this bacterium and motivated the
search for new strains. Later, strains with high activity were discovered
(Singer, 1977; Weiser, 1984; Wickremesingue and Mendis, 1980) that
led to the development of the use of Ls as a mosquito-control agent.
Insecticidal factors produced by Ls were identified in strains isolated
worldwide, and these isolates were classified according to their toxicity to
mosquitoes. Early studies showed that the high activity of some strains
was associated with the production of crystalline inclusions during the bac-
terial sporulation (Fig. 3.1) (de Barjac and Charles, 1983; Kalfon et al., 1984;
Payne and Davidson, 1984; Yousten and Davidson, 1982). The crystals are
synthesised during stage III of sporulation, and once formed, they remain
associated with the spore within the exosporium (Kalfon et al., 1984;
Yousten and Davidson, 1982). A study showed that mutant strains that were
blocked from the early stages of sporulation did not produce crystals and lost
their toxicity toward larvae, which confirmed the essential role played by the
crystals for the mosquitocidal activity of these strains (Charles et al., 1988).
The active crystals contain the binary (Bin) protoxin, which is the major
insecticidal protein produced by Ls (Baumann et al., 1985).
Figure 3.1 Micrography of Lysinibacillus sphaericus strain 2297 at the end of sporulation.
(A) Spore and crystal are in the left- and right side of the exosporium, respectively.
(B) Crystal lattice. Taken from Charles et al. (2010).
DNA homology group IIA, as will be described below. The highly toxic
strains isolated to date, can produce different insecticidal factors as the Bin
protoxin from crystalline inclusions and other toxins, such as the
mosquitocidal toxins (Mtxs) and the Cry48/Cry49 toxin, as will be
described in Section 2.
General classifications of Ls strains have been performed based on
different systems. The DNA homology of the strains (Krych et al., 1980)
and serotyping based on the flagellar antigen (de Barjac, 1990; de Barjac
Microbial Toxins for Mosquito Control 93
et al., 1985) are the most commonly used methods. Other approaches have
also been proposed to classify this diverse group, such as bacteriophage typ-
ing (Yousten, 1984b; Yousten et al., 1980), numerical classification using the
taxonomy of phenotypic features (Alexander and Priest, 1990), analysis of
the cellular fatty acids (Frachon et al., 1991), ribotyping (Aquino de
Muro et al., 1992), and the profiling of randomly amplified polymorphic
DNA (Woodburn et al., 1995). The toxic strains were found to belong
to few groups according to all the classifications used. DNA homology anal-
ysis placed the strains into five groups (IV) that probably represent distinct
species, and those that displayed some level of mosquito activity were clus-
tered in subgroup IIA. The flagellar antigen analysis generated approxi-
mately 49 serotypes, and nine serotypes host the highly toxic strains. The
serotype 5a5b contains active strains that produce the Bin toxin and includes
1593 and 2362, which are the most commonly strains used to date for the
production of larvicides. Other high-toxicity strains producing the Bin
toxin from DNA group IIA include the IAB881, IAB59, 2297 and
IAB872 strains, belonging to serotypes 3, 6, 25 and 48, respectively.
Table 3.1 shows some representative examples of strains and their classifica-
tion based on previously published data (Charles et al., 1996). Other highly
toxic strains have also been isolated and employed for the production of
biolarvicides. The C3-41 strain (serotype 5a5b) isolated in China, for
instance, has been extensively used in control programmes in that country
(Yuan et al., 1999). The screening of new strains for improved larvicidal
activity and optimal performance for large-scale production continues to
sustain the development of new products (Hire et al., 2010; Poophati
et al., 2013; Prabhu et al., 2013; Sun et al., 1996).
Table 3.1 Examples of Lysinibacillus sphaericus strains and their larvicidal properties to
mosquito larvae
Genes encoding
mosquitocidal proteinsb
Flagellar DNA Larvicidal Cry48/
Strain Origin serotype group activitya Crystal Mtxs Cry49
Kellen USA 1a IIA Low 1, 2, 3 Nd
K
SSII-1 India 2a2b IIA Medium 1, 2, 3 Nd
c
IAB881 Ghana 2a2b Nd Medium + 1, (2 and 3 +
Nd)
LP1-G China 3 IIA Medium + 1, (2 and 3 +
Nd)
1593M Indonesia 5a5b IIA High + 1, 2 (3 Nd)
2362 Nigeria 5a5b IIA High + 1, 2, 3
1691 El 5a5b IIA High + 1, 2, 3 Nd
Salvador
IAB59 Ghana 6 Nd High + 1, 2, 3 +
2297 Sri Lanka 25 IIA High + 1, 2, 3
IAB872 Ghana 48 Nd High + 1, (2 and 3 Nd
Nd)
C3-41 China 5a5b IIA High + 1 (pseudo)
2, 3, 4
a
Based on criteria defined by Charles et al. (1996).
+ Presence, absence.
b
c
Not determined.
Modified and extended from Charles et al. (1996)
larvicides. The control of Culex spp. has gained more importance recently
in regard to their role as vectors of emergent arboviruses, such as the West
Nile Virus that has provoked important epidemics in human populations
(Kramer et al., 2008; Petersen and Fischer, 2012). The control of anophe-
lines is a challenge, and previous studies have shown that some relevant spe-
cies involved in Plasmodium transmission are susceptible to Ls. The control
of Anopheles (An.) stephensi and An. sinensis in India and China, respectively,
showed the operational viability of Ls to control this group of mosquitoes
(Kumar et al., 1994; Thiery et al., 1996; Yuan et al., 2000). In addition to its
application for vector control in urban areas, the selective activity and bio-
compatibility of Ls is of great utility when the target species breed in envi-
ronmentally sensitive areas. The utilisation of Ls also raises concerns about
the selection of resistance. High levels of resistance achieved due to its
utilisation have been reported, and the major findings on this issue are pres-
ented in Section 5 along with the strategies that can be introduced for the
management of resistance in Section 6. These strategies can ensure the
effectiveness of Ls when used in the scope of integrated control programmes
and can overcome the potential onset of resistance. Different aspects of Ls
and its properties as an entomopathogenic bacterium have been covered by
previous reviews and book chapters that can provide additional information
(Baumann et al., 1991; Becker, 2000; Berry, 2012; Charles and Nielsen-
LeRoux, 2000; Charles et al., 1996, 2010; Delecluse et al., 2000; Lacey,
2007; Porter et al., 1993; Regis and Nielsen-LeRoux, 2000).
and Mtx toxins to larvae is, for instance, 3000-fold superior compared to the
SSII-1 strain that produces Mtx toxins only (Myers et al., 1979). Additional
studies on different strains have shown that Bin accounts for most activity
recorded for the sporulated cultures and this is the main active ingredient
of biolarvicides based on Ls, as reviewed by Charles et al. (1996). According
to the Insect Resistance Action Committee (www.irac-online.org), the
insecticidal toxins from Ls are classified into the mode of action group
11 (Moa11), along with Bti, and those agents are defined as bacterial
disruptors of insect midgut membranes. The midgut of mosquito larvae
is the central site for the action of these toxins, since they act following inges-
tion, are processed under specific conditions in this environment and they
act on specific receptors located on the epithelium, to cause mortality of lar-
vae. More details of insect midgut are provided in Chapter 1.
Mosquitoes are the principal targets of the Ls toxins and this is reflected in
the activity spectrum of the individual toxins. However, Ls toxicity to
Phlebotomus sandflies has been reported for high-toxicity strains 1593 and
2362 that may result in larval death and reduced fecundity of surviving
insects (Penner and Wilamowski, 1996; Robert et al., 1997, 1998;
Wahba, 2000). Strain 2362 also showed low toxicity against Lutzomyia
sandflies (Wermelinger et al., 2000). In addition, larvicidal effects of Ls
extracts against the nematode Trichostrongylus colubriformis have also been
reported (Bone and Tinelli, 1987) and some toxicity was seen against the
crustacean Palaemonetes pugio (Key and Scott, 1992).
Within the mosquitoes, there is differential toxicity to the species studied.
The most susceptible are Culex spp, in particular, those from the Cx. pipiens
complex, but one exception in this genus is Cx. cinereus larvae (Nicolas and
Dossou-Yovo, 1987). Anophelines including species of medical impor-
tance such as An. gambiae, An. stephensi, An. albimanus, An. quadrimaculatus,
An. darlingi and An. nuneztovari are also susceptible to the Bin toxin
(Arredondo-Jimenez et al., 1990; Davidson, 1989; Karch et al., 1992; Lacey
et al., 1988b; Rodrigues et al., 1998, 1999; Young et al., 1990). Aedes or
Ochlerotatus show a variable scenario including susceptible species such as
Oc. atropalpus, Ae. vexans and Oc. trivittatus, as well as Ae. aegypti larvae that
are refractory to Bin toxin (Berry et al., 1993; Delecluse et al., 2000;
Nielsen-Leroux and Charles, 1992). The lethal concentration (LC) of Ls
for these larvae is between 100- and 1000-fold higher than the respective
LC for Cx. pipiens larvae (Thiery and de Barjac, 1989). The screening of Ls
activity has also demonstrated susceptible larvae from Psorophora and Mansonia
species. On the other hand, Simulium larvae that are susceptible to Bti cannot be
Microbial Toxins for Mosquito Control 97
targeted by the Bin toxin. Table 3.2 presents a non-exhaustive list of mosquito
susceptibilities to Ls strains. The most common species targeted by Ls in field-
control trials or programmes are described in Section 4.
mechanisms that are still under investigation. Recently, activity of Bin toxins
against human cancer cells has also been reported (Luo et al., 2014).
The analysis of Bin proteins isolated from parasporal crystals was reported
in the mid-1980s (Baumann et al., 1985; Narasu and Gopinathan, 1986) and
initially it appeared that they were derived from a larger precursor protein
(Broadwell and Baumann, 1986). Subsequent cloning of the gene encoding
BinA (Berry and Hindley, 1987; Hindley and Berry, 1987) and BinB
(Baumann et al., 1987, 1988) showed that, in fact, the two components were
produced from a single operon as independent proteins of approximately
42 and 51 kDa, respectively. The proteins are produced in approximately
equimolar amounts and form a co-crystal in sporulating Ls whereas the indi-
vidual components expressed in recombinant Ls did not form crystals
(Charles et al., 1993). The combination of BinA + BinB forms crystals in
Ls and Bt strains but not in recombinant B. subtilis (Baumann and
Baumann, 1991; Broadwell et al., 1990a; Charles et al., 1993; Yuan
et al., 1999) suggesting that the former, insect pathogenic bacteria, encode
a factor that facilitates crystallisation and that is absent from B. subtilis. Bin
protein synthesis is enhanced by recombinant co-expression of the P20 pro-
tein from Bt (Park et al., 2007) but a region downstream of the bin operon in
Ls strain 2297 reduces Bin synthesis (Park et al., 2009). The activity of the
Bin toxin appears to be synergistic with the Cyt1Aa protein from Bt when
the two are co-expressed in acrystaliferous Bt (Li et al., 2000) but expression
of Cyt1Ab in Ls did not show synergy although it did help to overcome Bin
resistance in Culex larvae (Thiery et al., 1998). Other studies have shown
synergy of Bt Cyt and Cry toxins with Ls against wild-type or Bin-resistant
Culex, which may indicate synergy with Bin toxins although the use of Ls
cells in these assays may also indicate synergy with other toxins that they pro-
duce (see below) (Wirth et al., 2000a,c, 2001a, 2004).
Circular dichroism analysis has suggested that BinA and BinB are predom-
inantly composed of beta sheet (Hire et al., 2009; Kale et al., 2013;
Srisucharitpanit et al., 2012) although BinB in wild-type and truncated forms
has also been reported to contain considerable alpha helix (Tangsongcharoen
et al., 2011). Crystallization of BinB protein (Chiou et al., 1999;
Srisucharitpanit et al., 2013) and BinA/BinB co-crystals (Smith et al., 2004)
have been described and the structure of BinB has recently been published
(Srisucharitpanit et al., 2014). This protein has an N-terminal domain with
a beta-trefoil architecture found in lectins and a C-terminal region rich in
extended beta sheets that shows structural similarity to aerolysin beta pore for-
ming toxins. These results may suggest a role for the N-terminal region in
100 Maria Helena Neves Lobo Silva Filha et al.
receptor binding and roles for the C-terminal region and the related BinA
structure in formation of a beta pore. Association of the two proteins with
each other and the membrane may result in conformational changes to the
structures (Boonserm et al., 2006; Kale et al., 2013).
As described, the ingestion of the Bin proteins by mosquito larvae results
in their solubilisation in the alkaline environment of the gut (Charles, 1987)
and activation of the protoxin forms by proteolytic cleavage mediated by gut
proteinases (Aly et al., 1989; Broadwell and Baumann, 1987; Brownbridge
and Margalit, 1987; Davidson et al., 1987, 1990). These will activate both
subunits BinA (51 kDa) and BinB (42 kDa) into smaller polypeptides of
43 and 39 kDa, respectively. After proteolysis, the 43 kDa BinB derivative
results from the removal of 21 and 53 residues from the N- and C-termini,
respectively (Clark and Baumann, 1990). For the BinA active fragment of
39 kDa, cleavage of 10 and 17 amino acids occur in these respective posi-
tions (Broadwell et al., 1990c). The correct processing of the Bin subunits
and their presence in equimolar amounts are essential conditions that assure
the optimal activity of this Bin toxin (Broadwell et al., 1990b; Davidson
et al., 1990; Nicolas et al., 1993; Oei et al., 1990). Similar patterns of
protoxin cleavage occur on exposure to digestive enzymes from non-
susceptible larvae, indicating that the protoxin processing is not the origin
of insect specificity (Nicolas et al., 1990). In solution, NaOH solubilised
BinA/BinB crystal proteins may associate into a BinA2BinB2 heterotetramer
but this association may be lost on trypsin activation (Smith et al., 2005) with
activated proteins showing weak interactions between the two proteins
(Kale et al., 2013), although formation of oligomeric complexes between
activated toxins has also been suggested (Srisucharitpanit et al., 2012).
The solubilised and active toxin binds regionally to the larval midgut in
the gastric caecum and posterior midgut (Davidson, 1988, 1989; Mulla et al.,
1984a; Oei et al., 1992) leading to toxicity. In Culex larvae, the BinB com-
ponent of the toxin is responsible for receptor binding and the BinA com-
ponent subsequently binds to BinB or the BinB/receptor complex (Charles
et al., 1997; Oei et al., 1992). The situation in An. gambiae appears a little
more complex with a possible role for BinA in binding as well (Charles
et al., 1997). The receptor for Bin binding has been identified in Cx. pipiens
as a midgut-bound -glucosidase (Silva-Filha et al., 1999). Orthologs of this
protein have been identified in other mosquito species including An. gambiae
(Opota et al., 2008) and the refractory Ae. aegypti (Ferreira et al., 2010) and
differences in this receptor are believed to be the crucial factor in determin-
ing sensitivity (Section 3). Changes in the receptor are also known to cause
resistance to the Bin toxin (Section 5).
Microbial Toxins for Mosquito Control 101
Toxin binding to the midgut receptors is essential, and soon after the
ingestion of Bin crystals, cytopathological alterations can be observed.
The pathogenesis associated with Ls treatment was first described based
on the study on the action of strain SSII-1, which produces only Mtx toxins,
on Cx. quinquefasciatus larvae (Davidson, 1979). Subsequently, studies
showed the cytopathological alterations in midgut cells of larvae treated with
different strains all of which produced the Bin toxin as their major toxin
(Charles, 1987; de Melo et al., 2008; Silva Filha and Peixoto, 2003;
Singh and Gill, 1988). For Cx. pipiens, the major alterations observed in
the midgut cells are the intense disruption of microvilli, intense cytoplasmic
vacuolisations (or cytolysosomes) with broken membranes, pronounced
swelling of mitochondria and break-down of the endoplasmatic reticulum.
Ultra-structural effects investigated using an in vitro processed form of Bin
toxin to treat Cx. pipiens-cultured cells, showed similar effects as those seen
in midgut cells from larvae (Davidson and Titus, 1987). The study of Singh
and Gill (1988) also recorded damage in neural and muscles tissues that were
detected later than the major effects that were primarily observed in the mid-
gut cells. Treatment of resistant Cx. pipiens larvae that lacked the midgut
receptors with high concentration of Bin toxin showed minor alterations
that were comparable with those observed for Ae. aegypti, a refractory species
that does not have functional receptors in their midgut (Charles, 1987; de
Melo et al., 2008). Physiological studies of the Bin action are not available
except for one that shows an inhibition of the oxygen uptake of mitochon-
dria and in the activity of the enzyme choline acetyl transferase in larvae
treated with the Bin toxin (Narasu and Gopinathan, 1988).
The mode of action of the Bin toxin, following receptor binding, remains
somewhat unclear. Many reports have suggested that to exhibit toxicity, both
BinA and BinB components are absolutely required (Broadwell et al., 1990b;
Charles et al., 1993; Nicolas et al., 1993; Oei et al., 1990), with optimal activity
reported when components are present in approximately equimolar amounts
(Davidson et al., 1990). Nevertheless, toxicity of BinA in purified form (Hire
et al., 2009) or produced in B. subtilis (de la Torre et al., 1989) or in Bt (Nicolas
et al., 1993) has been reported by some authors. When the Cx. pipiens Bin
receptor Cpm1 was expressed in Madin Darby canine kidney cells, patch-
clamp experiments showed that toxin binding is followed by the induction
of currents that are likely to be due to the opening of pores (Pauchet et al.,
2005). Experiments with Culex cells in culture (Cokmus et al., 1997) and arti-
ficial membranes (Schwartz et al., 2001), also suggest that the toxin may be able
to form pores and indicated that BinA was better able to form pores than BinB,
consistent with the model whereby BinB is the receptor-binding component
102 Maria Helena Neves Lobo Silva Filha et al.
and BinA forms a pore. In contrast, a separate report described the ability of
BinB to interact with artificial membranes and form pores in the absence of
BinA (Boonserm et al., 2006). This pore formation was proposed to be
through membrane insertion of beta sheet rather than alpha helical structures.
Thus, an alternative model for Bin toxicity may involve receptor binding and
pore formation by BinB coupled with an unknown role for BinA or by a
BinA/BinB complex and this may be supported by the crystal structure of
BinB (Srisucharitpanit et al., 2014).
In addition to the possibility of pore formation, a further effect is char-
acteristic of Bin intoxication. The vacuolisation of target cells is seem
(Charles, 1987; Davidson, 1988; Pauchet et al., 2005) accompanied by
the uptake of labelled toxins into vesicles (Davidson, 1988), a phenomenon
that only occurs when both BinA and BinB components are present
together (Oei et al., 1992). A detailed study of this phenomenon was carried
out using Madin Darby cells expressing the Bin receptor Cpm1 (Opota
et al., 2011). This investigation showed the opening of cationic pores in
the membrane and demonstrated that the large vacuoles formed in target
cells were autophagic. These structures were transient but, having dis-
appeared from the cells, these vacuoles then reappeared following cell divi-
sion: a novel phenomenon termed post-mitotic vacuolation. The uptake of
Bin into the cells, along with their receptor, was shown to be via recycling
endosomes; structures that are distinct from the large transient autophagic
vesicles. Thus, Bin intoxication induces autophagy, while Bin uptake in sep-
arate structures protects it from degradation by targeting to recycling path-
ways. The overall significance of these events for toxicity remains to be
clarified but Bin trafficking may allow it access to tissues beyond the midgut.
The BinA and BinB proteins are related to each other and to a family of
Bin-like proteins including Cry49 from Ls (see below), Cry35 and Cry36 from
Bt, and sequences of unknown function from B. cereus group strains (e.g. acces-
sion number ZP_17404242) and Chlorobium phaeobacteroides (accession num-
ber Y_911930) (Baumann et al., 1988; Jones et al., 2007). Cry36 acts alone to
cause insect mortality, whereas Cry35 requires the 14 kDa Cry34 protein for
toxicity and Cry49 requires the three-domain family toxin Cry48 for its func-
tion. The various interactions that this family of proteins may require for
toxicity, further complicates our understanding of their modes of action.
The Bin toxin proteins themselves are highly conserved. Strains isolated
from around the world produce Bin toxins (Priest et al., 1997) but only six
variants of BinA (in which nine amino acids are altered) and four variants of
BinB (in which six amino acids are altered) have been described (Hire et al.,
2009; Humphreys and Berry, 1998; Priest et al., 1997). Two variants, Bin1
Microbial Toxins for Mosquito Control 103
and Bin2 have been shown to share the same receptor (Silva-Filha et al.,
2004) and cross-resistance between variants is seen (Yuan et al., 2003). Nev-
ertheless, Bin toxin variants can show differential activity against mosquito
targets. BinA variants were shown to alter the activity against the marginal
target Ae. aegypti and to alter the progression of growth and mortality for Cx.
quinquefasciatus and these effects were localised to amino acids 99 and 104 in
this protein (Berry et al., 1993). Reciprocal exchange of the amino acid at
position 93 of the BinA protein between the BinA2 variant, which is highly
active against Cx. pipiens larvae and the BinA4 variant, which is non-toxic to
this insect, showed that this residue was also a key determinant of activity
(Yuan et al., 2001). Deletion experiments have defined the essential core
regions of the Bin toxins. BinA can be truncated by 17 residues at both
the N- and C-termini without loss of toxicity. BinB can be truncated by
34 residues at the N-terminus and 53 residues at the C-terminus without
loss of toxicity (Broadwell et al., 1990c; Clark and Baumann, 1990, 1991;
Limpanawat et al., 2009; Oei et al., 1990; Sebo et al., 1990). Analysis of
the binding of non-toxic variants suggested that the N-terminal region of
BinB may have a role in interaction with the receptor and its
C-terminus, along with both the N- and C-termini of BinA may be
involved in the interaction of the two proteins (Oei et al., 1992). Predictions
of structural disorder within the BinA and BinB proteins have suggested that
the N- and C-termini may be flexible, consistent with a role in protein
protein interactions (Kale et al., 2013). More detailed analysis of the
N-terminal region of BinB and receptor binding, confirmed the importance
of residues 33158 in this interaction and, particularly, the sequences Ile-
Arg-Phe (residues 8587) and Phe-Gln-Phe (residues 147149) (Romao
et al., 2011). Mutagenesis studies on BinB indicated that individual substi-
tution of Pro35, Glu36, Phe41 and Tyr42, resulted in reduced activity but all
were able to interact with BinA (Singkhamanan et al., 2013). Pro35 and
Phe41 to alanine substitutions could bind to the larval midgut at a compa-
rable level to the wild-type BinB but binding of the Tyr42 to alanine mutant
was reduced and the Pro35Ala replacement decreased penetration of the
membrane. Block mutation of BinB from residues 113150 showed the
protein to have some tolerance to mutation in this region (Singkhamanan
et al., 2010). Alanine replacement of individual amino acids Phe149 and
Tyr150 resulted in loss of toxicity and loss of midgut binding for the latter
mutant. Toxicity could be rescued by replacement of these two residues
with other aromatic residues. In BinB, the substitution of Cys67 or
Cys161 reduced BinB interaction with BinA and eliminated toxicity while
replacement of Cys241 had no effect (Boonyos et al., 2010). In similar
104 Maria Helena Neves Lobo Silva Filha et al.
2.3. Cry48/Cry49
While the Bin toxin is the major sporulation-associated toxin in most Ls
strains, the ability of the spores of a small number of strains to overcome
Bin resistance in mosquitoes, indicated the presence of another toxin in
these strains and an approximately 49 kDa protein was identified as a candi-
date (Pei et al., 2002; Shi et al., 2001). The genes encoding the Cry49 pro-
tein and a second crystal protein (Cry48) related to the three-domain toxins
of Bt has been cloned and expressed ( Jones et al., 2007) and show a very
narrow target range, so far active only against Culex mosquitoes ( Jones
et al., 2008). It is of interest that the two proteins form a novel type of
Bin toxin since both components are necessary for activity and no other
three-domain protein has ever been shown to have a requirement for
another protein for its activity ( Jones et al., 2007). Consistent with the pres-
ence of a Bin-type toxin and a three-domain toxin, the cytopathology of
Culex cells exposed to the Cry48/Cry49 toxin shows features of both toxin
types, including the vacuolation observed on Bin intoxication (de Melo
et al., 2009) (Fig. 3.2). Similar effects were seen when Bin susceptible larvae
were treated with the synergistic combination of Bin and Cry11Aa toxins.
1991). The toxin is present in many high- and low-toxicity strains but is lac-
king in some strains that produce Bin toxin such as LP1-G (Liu et al., 1993;
Priest et al., 1997) and is present as a disrupted pseudogene in others (e.g.
strain C3-41) (Hu et al., 2008). The toxin is active against both Cx.
quinquefasciatus and Ae. aegypti larvae but has no effect on the mosquito
Toxorhynchites splendens (Thanabalu, 1992; Thanabalu et al., 1992a). It also
has low-level toxicity against Chironomus riparus (Partridge and Berry,
2002). Mtx1 acts synergistically with Mtx2 from Ls (Rungrod et al.,
2009) and Cry11 but is antagonistic with Cyt1Aa (Wirth et al., 2007). In
combination with other toxins from mosquitocidal Bt, synergy can also
be seen and mosquitoes resistant to individual Bt toxins may show some
cross-resistance to Mtx1 (Wirth et al., 2014).
The gene encoding Mtx1 is preceded by an inverted repeat region, char-
acteristic of a binding site for a regulatory protein (Thanabalu, 1992). The
production of a reporter protein driven by the mtx1 promoter was found to
be higher in B. subtilis than in Ls, suggesting regulation in the latter species
(Ahmed et al., 1995) and it has been speculated that the gene upstream of
mtx1 may encode a BglG family regulator of Mtx1 production (Berry,
2012). The toxicity of purified Mtx1 is high, showing equivalent potency
to the Bin toxin (Thanabalu et al., 1992a) and the low-level toxicity of
Ls strain SSII-1 is thought to be due to low levels of Mtx1 production
and to the degradation of Mtx1 by the producing bacterium (Thanabalu
and Porter, 1995). The proteinase responsible for this degradation has been
identified (Wati et al., 1997; Yang et al., 2007b) and is known as sphericase/
sfericase (Almog et al., 2003; Yoshida et al., 1977). Some success in
protecting the toxin by expression in proteinase negative strains of Ls has
been reported (Thanabalu and Porter, 1995) but the broad specificity of
sfericase has blocked attempts to stabilise the toxin by mutagenesis (Yang
et al., 2007b).
Produced as a 100 kDa protein, Mtx1 is processed by gut enzymes to an
approximately 27 kDa moiety, with regional sequence similarity to ADP-
ribosylating toxins and a C-terminal, 70 kDa moiety containing lectin-like
beta-trefoil repeat sequences (Hazes and Read, 1995; Thanabalu et al.,
1992a). This cleavage produces the two subunits that enable Mtx1 to func-
tion as a two component, AB toxin in which the 70 kDa protein is expected
to act as the receptor-binding unit, probably mediated by the lectin-like
motifs. This is expected to allow the enzymatic subunit (27 kDa) to enter
the susceptible cells where it will modify target proteins by ADP-
ribosylation. The C-terminal 70 kDa protein is able to cause morphological
Microbial Toxins for Mosquito Control 107
changes in Cx. quinquefasciatus and Ae. aegypti cells in culture but not to An.
gambiae cells or human HeLa cells (Thanabalu et al., 1993). However, both
components are required for toxicity to insects. The ADP-ribosyl transferase
subunit has been shown to promote ADP-ribosylation of Mtx1 components
and proteins of 38 and 42 kDa from Cx. quinquefasciatus cell extracts
(Thanabalu et al., 1993). Following activation with chymotrypsin, the
kinetics of the Mtx1 ADP-ribosyl transferase have been elucidated with
respect to the NAD+ substrate using soybean trypsin inhibitor as an artificial
substrate and ribosylation is seen to occur at arginine residues (Carpusca
et al., 2006; Schirmer et al., 2002a). The ADP-ribosylation activity of the
catalytic subunit can be inhibited by the 70 kDa protein, which binds
non-covalently, mediated by residues Asp273 and Asp275, and a
C-terminal region (Carpusca et al., 2004). Furthermore, transfection of
the catalytically active subunit into HeLa cells produced cytotoxic effects
(Schirmer et al., 2002a), suggesting that specificity of Mtx1 is due to the
70 kDa component. Expression of the enzymatic subunit alone in Escherichia
coli proved problematic due to toxicity of the protein to host cells. This was
shown to be due to ADP-ribosylation of the bacterial elongation factor Tu
by the toxin (Schirmer et al., 2002b).
The structure of the catalytic subunit (with a small sequence from the
N-terminus of the 70 kDa protein) has been solved (Reinert et al., 2006)
as has the structure of the full-length protoxin (Treiber et al., 2008). In
the latter form, the putative receptor-binding moiety (70 kDa subunit)
can be seen with the four lectin domains curled around the catalytic domain
and the authors propose a mechanism for the toxin whereby proteolytic acti-
vation occurs at the exposed site between the two subunits; the binding sub-
unit then interacts with glycolipid via the lectin-like domains and is
endocytosed with the catalytic unit non-covalently attached. At low pH,
the Mtx1 forms a heptamer, membrane insertion of N-terminal segments
occurs, leading to translocation of the catalytic core into the cytosol. This
translocation also separates the enzyme from the 70 kDa protein, thereby
freeing it from inhibition.
function to Mtx1 but is related to Mtx3, later identified in SSII-1 (Liu et al.,
1996), and Mtx4 identified in the genome of the high-toxicity strain C3-41
(Hu et al., 2008), a strain that also encodes Mtx2 and Mtx3 along with a
pseudogene related to this family of proteins, indicating probable gene
duplication in this family of genes (Berry, 2012). In addition, the proteins
are related to Clostridium epsilon toxin and Cry15, Cry23, Cry33 and
Cry38 from Bt (de Maagd et al., 2003) in the Pfam ETX_MTX2 superfamily
of proteins. The similarity to epsilon toxin suggests that these Mtx toxins act
by pore formation. Mtx3 is active against Cx. quinquefasciatus larvae and is
highly conserved, with only one conservative amino acid replacement
reported (Liu et al., 1996). Most studies on this family have focused on
Mtx2. This protein shows less conservation than Mtx3. Although a series
of individual mutations in the mtx2 genes did not alter their toxicity to
Cx. quinquefasciatus, one mutant, from lysine to threonine at residue 224,
abolished activity. Double mutants at positions 224 and 279 significantly
affected activity against this insect and Ae. aegypti so that specificity can,
effectively, be switched between the two species (Chan et al., 1996). The
Mtx2 family of proteins features a putative N-terminal signal sequence,
although their secretion has not been confirmed. Mtx2 can be truncated
by up to 23 amino acids at its N-terminus (including this signal sequence)
without loss of activity but deletion of only five residues from the
C-terminus produced inactive toxin (Phannachet et al., 2010). Mtx2 shows
instability on exposure to conditioned media or the Ls proteinase sfericase
(Yang et al., 2007b) so that, like Mtx1, it appears to be a rather short-lived
toxin. Mtx2 is synergistic with Cyt1Aa and Cry11Aa (Wirth et al., 2007)
while interactions with other Bt toxins is somewhat complex with some
cross-resistance exhibited (Wirth et al., 2014).
2.6. Sphaericolysin
Sphaericolysin is a cholesterol dependent cytolysin that was discovered in Ls
strain A3-2 that was isolated from the crop of Myrmeleon bore (ant lion) larvae
(Nishiwaki et al., 2007). The protein shows injection toxicity against the
cockroach, Blattella germanica, and, to a lesser extent, against the common
cutworm, Spodoptera litura. Sphaericolysin induces pores in erythrocyte
membranes and inserts to form a complex with an external diameter of
around 35 nm. The residue Tyr187, equivalent to Tyr159 in perfringolysin
is important in pore formation and the Tyr to alanine mutant markedly
reduced haemolytic activity as did dosing with cholesterol. A cholesterol
binding motif is located close to the C-terminus of the protein and is
Microbial Toxins for Mosquito Control 109
Figure 3.3 Alignment of sphaericolysin and related proteins. Identical residues are
shaded green or cyan, spaces introduced to optimise the alignment are shaded grey.
The sequences shown are Ls A3-2 (sphaericolysin from Lysinibacillus sphaericus strain
A3-2, accession number BAF62176), Ls B354 (sphaericolysin from L. sphaericus strain
B354, accession number EU043116), Bcer (anthrolysin from Bacillus cereus accession
number ZP_03231124), BtkAlv (alveolysin from B. thuringiensis subsp. kurstaki accession
number ZP_04117355), P alv (P. alvei alveolysin accession number ZP_10866820),
B14905 (perfringolysin O precursor from Bacillus sp B14905 accession number
ZP_01724867). Alignment generated using ClustalX (Thompson et al., 1994).
112 Maria Helena Neves Lobo Silva Filha et al.
Section 2, the activated form of the toxin recognises and binds to the midgut
epithelium of susceptible larvae. The cytopathological effects and the lethal
action observed in Bin-treated larvae depend, essentially, on the ability of
this toxin to bind to the midgut cells. Treatment of larvae with labelled-
toxin and further analysis of the midgut showed the interaction of the
Bin toxin along the epithelium, and different patterns were described
(Davidson, 1988, 1989; Oei et al., 1992). For Cx. pipiens, the most suscep-
tible species, Bin displays a marked and regionalised binding to the gastric
caeca and posterior midgut (Fig. 3.5). For anopheline larvae such as An.
gambiae, An. stephensi, An. albimanus and An. quadrimaculatus, the binding
to their midguts was variable and less defined than that found for Cx. pipiens.
For Ae. aegypti, Bin refractory larvae, this interaction was very weak, com-
pared to the Bin binding for the other species. The midgut binding patterns
observed are directly correlated to the in vivo susceptibility for the species
studied.
Quantitative assays that were performed to measure the Bin binding
affinity for the proteins from the midgut, confirmed this association.
In vitro binding assays between radiolabelled toxin and midgut brush-border
membrane fractions (BBMF) enriched with proteins from the microvilli of
apical cells, showed that Bin toxin binds with high affinity to BBMF from
Cx. pipiens and Cx. quinquefasciatus larvae (Nielsen-Leroux and Charles,
1992; Nielsen-Leroux et al., 1995, 1997, 2002; Oliveira et al., 2004;
Silva-Filha et al., 1997, 2004, 2008). Binding is saturable, reversible and
Figure 3.5 Binding of fluorescent-labelled binary toxin regionalized in the gastric caeca
(GC) and posterior midgut (PMG) of Culex quinquefasciatus larvae. Figure kindly provided
by Colin Berry, Cardiff University (UK).
114 Maria Helena Neves Lobo Silva Filha et al.
3.2. Receptors
Midgut-bound proteins from insect larvae play an essential role as receptors
for insecticidal proteins from entomopathogenic bacteria. Aminopeptidases
(APNs), cadherins (CADs), alkaline phosphatases (ALPs) and -amylase
have been characterised as receptors to the Cry toxins from Bt strains in
lepidopteran and mosquito larvae (Abdullah et al., 2006; Chen et al.,
2009a,b; Fernandez et al., 2006; Fernandez-Luna et al., 2010; Hua et al.,
2008, 2009; Jurat-Fuentes and Adang, 2004; Knight et al., 1994; Luo
et al., 1997; Vadlamudi et al., 1995; Valaitis et al., 2001). More information
on receptors of Bt toxins is presented in Chapter 2. The receptors of
the Bin toxin from Ls are -glucosidases and they were characterised
in three species as follows: Cpm1 for Cx. pipiens maltase 1 (Darboux
et al., 2001; Silva-Filha et al., 1999), Cqm1 for Cx. quinquefasciatus maltase
1 (Romao et al., 2006) and Agm3 for An. gambiae maltase 3 (Opota et al.,
2008). Aedes aegypti has also a gene that encodes an orthologue of the Bin
toxin receptor, Aam1 protein (Ae. aegypti maltase 1) that displays 74% iden-
tity to the Cqm1 -glucosidase; however, Aam1 is not able to bind to Bin
toxin (Ferreira et al., 2010; Nene et al., 2007; Opota et al., 2008; Romao
et al., 2006).
Microbial Toxins for Mosquito Control 115
7
Bound 125l-toxin (pmol / mg)
Specific
6
4
Non-specific
3
0
0 50 100 150 200
125
[Free l -toxin] (nM)
7
Bound 125l-toxin (pmol / mg)
3
2
0
0 50 100 150 200
[Free 125l -toxin] (nM)
Figure 3.6 In vitro saturation binding assays between binary toxin and brush-border
membrane fractions (BBMF) from fourth instar Culex quinquefasciatus larvae from a sus-
ceptible colony (A) and from CqRL1/2362, a Lysinibacillus sphaericus-resistant colony (B).
Taken from Oliveira et al. (2004).
Table 3.3 Mosquito susceptibility to Lysinibacillus sphaericus and the capacity of the Binary (Bin) toxin to interact with larval midgut
Characteristic Culex pipiens a Cx. quinquefasciatus a Anopheles gambiae b An. stephensi b Aedes aegypti c
The expression of Cpm1 in Madin and Darby kidney canine (MDKC) cells
was also successfully achieved. Both catalytic activity and Bin binding capac-
ity were observed, in addition, this recombinant receptor was able to medi-
ate the cytopathological effects derived from the treatment with Bin toxin
(Pauchet et al., 2005). The expression of Cpm1/Cqm1 as membrane-bound
proteins is essential for the activity of Bin toxin on the larva and they are
bound to the epithelium through a GPI anchor. These proteins can be sol-
ubilized from the midgut through the action of the enzyme pho-
sphatidylinositol phospholipase C that specifically releases GPI-bound
proteins from the cell surface (Darboux et al., 2002; Ferreira et al., 2010;
Silva-Filha et al., 1999). Mutations in the cpm1/cqm1 genes that prevent
Microbial Toxins for Mosquito Control 119
expressed in a comparable level to that observed for Cqm1 (Fig. 3.7), how-
ever, this protein does not display capacity to bind the Bin toxin (Ferreira
et al., 2010). One major difference between Cqm1 and Aam1 is
N-glycosylation and, although both sequences present predicted sites
(Table 3.4), only Aam1 was found to have carbohydrates added that signif-
icantly increase its size, while the glycosylation of Cqm1, if present, could
not be detected (Ferreira et al., 2010). Cqm1 treatment with the endo-
glycosidase PNGase F does not abolish its capacity to bind to BinB,
suggesting that this interaction does not involve glycans. On the other hand
the hypothesis that glycans from Aam1 could hide or prevent the access of
the Bin toxin to the binding site was refuted because the removal of carbo-
hydrates by deglycosylation with PNGase F did not provoke any effect in
this aspect (Ferreira et al., 2010). The differential role of this major post-
translational modification in Aam1 remains unknown. Further investigation
using wild and mutant Cqm1 and Aam1 proteins expressed in Sf9 cells
showed that non-conserved residues of Cqm1 are needed for the BinB sub-
unit binding (Ferreira et al., 2014). This study showed that a segment of
Cqm1 N-terminus is responsible for binding to BinB subunit and the
replacement of the doublet 159GG160, located in a loop region, by the
respective residues (KL) from Aam1 protein abolished binding, showing that
they are required for this interaction. This loop from the Aam1 ortholog
Microbial Toxins for Mosquito Control 121
shows other divergent residues and the insertion of five amino acids that
could be implicated in the lack of capacity of this protein to bind BinB.
Those residues showed to be essential for binding, probably by making pos-
sible the access of Bin toxin to amino acids that may act as binding sites
themselves. Additional investigation of these proteins will certainly be
important to understand the molecular basis of Bin toxin selectivity in the
insect midgut.
The studies of the molecules involved in the mode of action and selec-
tivity of the Bin toxin has been restricted to the midgut receptors. Despite
the role of -glucosidases as the major target site of the Bin toxin, it is pos-
sible that novel molecules and pathways could be involved in the mode of
action and selectivity of Ls, as demonstrated by recent studies in the action of
Cry toxins to mosquitoes and lepidopterans. The immune defence involving
the mitogen-activated protein kinase p38 (MAPK p38) pathway seems to
participate in the response to pore-forming toxins and it is responsible for
protecting Ae. aegypti larvae from the action of Cry11Aa toxin (Cancino-
Rodezno et al., 2010; Porta et al., 2011). Silencing of genes from this path-
way rendered larvae hypersensitive to that toxin indicating its importance to
modulate the Cry toxin action. Mutations in the ABC transporters were
directly linked to the resistance of lepidopteran larvae to Bt Cry1A toxins
also proving their involvement in the mode of action (Atsumi et al.,
2012; Gahan et al., 2010). The role of other molecules in the mode of action
of Bin toxin, therefore, remains a field to be explored.
Table 3.5 Field trials using Lysinibacillus sphaericus-based products to control mosquito
larvae
Targeted species Environment References
Culex quinquefasciatus Containers Kuppusamy et al. (1987),
Lacey et al. (1984)
Sod-lined potholes Lacey et al. (1988b)
Polluted sites, polluted Hougard (1990), Karch et al.
ponds, cesspools, cesspits, (1991, 1990), Kumar et al.
catch basins, settling basins (1996), Mulla et al. (1997,
1999), Nicolas et al. (1987),
Paing et al. (1987), Siegel
and Novak (1997, 1999),
Skovmand and Bauduin
(1997), Skovmand and
Sanogo (1999), Su and
Mulla (1999), Su (2008)
Diverse urban habitats Andrade et al. (2007),
Barbazan et al. (1997),
Hougard et al. (1993), Mulla
et al. (2001), Regis et al.
(1995, 1996, 2000b), Silva-
Filha et al. (2001),
Skovmand et al. (2009)
Dairy wastewater Jones et al. (1990), Mulla
et al. (1984a)
Cx. quinquefasciatus, Cx. Polluted pools Barbazan et al. (1998),
annulirostris, Anopheles Brown et al. (2004),
funestus, An. albimanus Montero Lago et al. (1991),
Ragoonanansingh et al.
(1992)
Cx. quinquefasciatus, Cx. Experimental containers Su and Mulla (1999)
stigmatosoma, Cx. tarsalis
Cx. quinquefasciatus, Cx. Diverse urban habitats Yadav et al. (1997)
tritaeniorhynchus, Cx.
vishnui
Cx. pipiens Artificial pools, settling Berry et al. (1987), Karch
basins et al. (1990)
Cx. pipiens, Cx. restuans Catch basins Siegel and Novak (1997)
Continued
124 Maria Helena Neves Lobo Silva Filha et al.
Table 3.5 Field trials using Lysinibacillus sphaericus-based products to control mosquito
larvaecont'd
Targeted species Environment References
Cx. pipiens, Ochlerotatus Waste water lagoons Berry et al. (1987)
trivittatus,
Cx. peus, Cx. Dairy wastewater Matanmi et al. (1990), Mulla
stigmatosoma et al. (1988a)
Cx. nigripalpus Wastewater Lacey et al. (1988a)
Cx. restuans Woodland pools Lacey et al. (1988b)
Culex spp Sod-line ponds Lord (1991)
An. gambiae Irrigation ponds, sunlit Fillinger et al. (2003), Karch
ponds, rain puddles et al. (1991), Skovmand and
Bauduin (1997), Skovmand
and Sanogo (1999)
An. stephensis Artificial breeding sites Kumar et al. (1994)
An. arabiensis Pools, rice fields water Romi et al. (2003), Shililu
ditches et al. (2003)
An. quadrimaculatus, Rice fields Dennett et al. (2001),
Psorophora columbiae, An. Dennett and Meisch (2000),
crucians Lacey et al. (1986, 1988b)
An. albimanus Lake habitats Rivera et al. (1997)
An. albimanus, Culex spp Experimental potholes Arredondo-Jimenez et al.
(1990)
An. culicifacies Rice fields Sundararaj and Reuben
(1991)
An. darlingi Artificial breeding sites Berrocal et al. (2000),
Galardo et al. (2013),
Rodrigues et al. (2008)
An. nuneztovari Field pools Rojas et al. (2001)
An. marajoara, An. Field pools Moreno et al. (2010)
triannulatus, An.
braziliensis
An. aquasalis Field ponds Berti and Gonzales (2004),
Moser et al. (2002, 2012)
Mansonia spp, Ma. Rice ditches and rice Floore and Wardz (2009),
indiana, Ma. uniformis, ponds Pradeepkumar et al. (1988),
Ma. dyari Yap et al. (1991)
Microbial Toxins for Mosquito Control 125
Table 3.5 Field trials using Lysinibacillus sphaericus-based products to control mosquito
larvaecont'd
Targeted species Environment References
Ps. columbiae Rice field plots, flood- Bowles et al. (1990), Groves
water habitats and Meisch (1996), Lacey
et al. (1986), Lord (1991)
Ps. columbiae, Oc. Irrigated fields Mulla et al. (1985), Mulla
nigromaculis et al. (1988b)
Oc. nigromaculis Irrigated pastures Mulla et al. (1988b)
Oc. triseriatus, Cx. Tires Siegel and Novak (1997,
restuans, Cx. pipiens 1999), Siegel et al. (2001)
Aedes vexans Riparian woodland Becker (2003)
Culiseta incidens Tires Kramer (1990)
of Ls-based larvicides, and there are successful examples of trials for control-
ling these mosquitoes in their typical habitats, which are commonly situated
in rural areas, such as irrigated fields, pastures, rice fields and other flooded
environments (Table 3.5). Numerous trials to control mosquitoes based on
Ls larvicides have been conducted; however, it is not possible to establish a
common methodology to employ this agent, and each programme should be
designed according to the targets biology and ecology, the biotic and abiotic
conditions, and other factors that play a role in the proliferation of that
species.
Ls is less susceptible to this factor than Bti (Silapanuntakul et al., 1983), both
produce insecticidal proteins and, thus, can suffer degradation due to solar
radiation (Burke et al., 1983; Karch and Charles, 1987). Prolonged Ls per-
sistence is often observed in septic tanks and other breeding sites that are
protected from the sunlight (Silva-Filha et al., 2001). In a similar manner,
Bti can achieve enhanced performance under shadowed conditions
(Araujo et al., 2007; Melo-Santos et al., 2009). Sedimentation of the
crystal-spore complex also reduces the activity of Ls because Culex and
Anopheles larvae feed at the surface of the breeding sites. Some studies
showed a correlation between the crystal spores settling and the loss of tox-
icity in experimental sites (Davidson et al., 1984; Lacey and Lacey, 1990;
Skovmand and Guillet, 2000). Water flow, typically found in breeding sites
such as catch basins or water ditches in urban areas, has a negative impact on
the persistence of Ls because the current carries away the active ingredient.
Ls persistence for long periods is most likely the major feature that con-
tributes to its field effectiveness, particularly in urban areas (Davidson et al.,
1984; Mulla et al., 1984a; Mulligan et al., 1980; Silapanuntakul et al., 1983).
This persistence occurs because Ls has the ability to recycle in mosquito
cadavers or in the soil of breeding sites. This process involves a new cycle
of spore germination, vegetative development and the production of new
batches of spores and crystals that can be released in the environment and
sustain the larvicidal activity. The recycling has been demonstrated under
laboratory and field conditions and can take place in some mosquito species
(Becker et al., 1995; Correa and Yousten, 1995; Des Rochers and Garcia,
1984; Hertlein et al., 1979; Karch and Coz, 1986; Menon et al., 1982;
Nicolas et al., 1987). Laboratory studies have shown that de novo production
of spores is approximately 20-fold greater than the amount that was origi-
nally ingested by larvae (Charles and Nicolas, 1986). Recycling can provide
effective results if the favourable conditions required for bacterial germina-
tion occur (mainly in the mosquito cadavers) and if the newly produced
spore crystals are located in the feeding zone of larvae. The availability of
spore crystals on the bottom of breeding sites cannot contribute to activity
against larvae (Singer, 1980).
to develop a strategy for large-scale field trials aiming to evaluate the efficacy
of Ls against Culex populations under different ecological conditions in
Brazil, Cameroun, India, Sri Lanka and Tanzania (WHO, 1993). The
projects were conducted in four steps, according to the protocol developed
by TDR, which also considered the effects of vector control on the trans-
mission of Wuchereria bancrofti: a baseline collection of data (i), a preparation
phase (ii), the implementation of an 18 month-larviciding period (iii), and a
6-month follow-up phase after larviciding stopped (iv). Entomological and
parasitological data gathered from all the phases showed a remarkable impact
of Ls treatment of breeding sites on reducing the vector-biting rate and gen-
erating a significant decline in human exposure to filarial infective larvae
(Barbazan et al., 1997; Becker, 2000; Maxwell et al., 1999; Regis et al.,
2000b). The output of these projects, together with previous knowledge
and experience from field trials, contributed to constructing a solid basis
for the design of large-scale implementation of Ls-based larvicides under dif-
ferent ecological conditions. The large-scale use of Ls also resulted in the
emergence of resistance in larvae of the Cx. pipiens complex (see
Section 5). However, given the known advantages of Ls over conventional
larvicides, that discovery promptly led to a research effort to elucidate the
resistance process and mechanisms. As a consequence, solutions to avoid,
delay or manage resistance were soon made available (see Section 6). Since
then, Ls has been applied in association with Bti in rotation or as a mixture,
or associated with other control agents, and the susceptibility of mosquito
larvae to the Bin toxin should be duly monitored.
Table 3.6 Examples of field trials using combined Lysinibacillus sphaericus and Bacillus
thuringiensis serovar. israelensis strategies for mosquito control
Country Targeted species References
Kenya, Mbita Anopheles gambiae, An. funestus Fillinger and Lindsay
(2006)
Gambia, Farafeni An. gambiae, An. arabiensis, An. melas Majambere et al.
Town (2007)
Turkey, Antalya Culex pipiens Cetin et al. (2007)
Colombia, Cali Cx. quinquefasciatus, Aedes aegypti Giraldo-Calderon
et al. (2008)
Cote dIvoire, An. funestus, An. gambiae Tchicaya et al. (2009)
Tiemelekro
Tanzania, Dar es An. gambiae, An. funestus, An. coustani, Geissbuhler et al.
Salaam Cx. quinquefasciatus (2009)
Kenya, Highland An. gambiae, An. funestus, An. arabiensis Fillinger et al. (2009)
Valley
Poland, Wroclaw Anopheles spp. Rydzanicz et al.
(2009)
Benin Cx. quinquefasciatus and other Lingenfelser et al.
mosquito species (2010)
Kenya, Malindi An. gambiae, Cx. quinquefasciatus Mwangangi et al.
(2011)
USA, Stratford Cx. pipiens, Cx. restuans, Ae. japonicus Anderson et al. (2011)
CN
USA, California Cx. tarsalis, Ochlerotatus melanimon Dritz et al. (2011)
Switzerland, Cx. pipiens, Ae. albopictus Guidi et al. (2013a)
Chiasso, Ticino
Tanzania, Dar es An. gambiae, An. funestus Maheu-Giroux and
Salaam Castro (2013)
Brazil, Amapa An. darlingi Galardo et al. (2013)
countries (Table 3.6). A common finding in these field trials is that all the
tested species of malaria vectors were highly susceptible to Ls, in each loca-
tion that testing was conducted. Furthermore, an evaluation of the impact of
the biolarvicide on the vector population and malaria transmission was per-
formed in most of these works. As an example, Fillinger and Lindsay (2006)
Microbial Toxins for Mosquito Control 129
have demonstrated that microbial larvicides (Bti and Ls) reduced the malarial
vector mosquito larvae and adult females by more than 90% in West Kenya.
In a location in Tanzania, one year of community-based larvicide applica-
tion reduced transmission by the primary malaria vector, An. gambiae, by
31% (Fillinger et al., 2008). As described, the first field trials of Ls against
mosquitoes were conducted, mainly in Africa, against Culex and Anopheles
species in the 1980s (Table 3.5). Currently, there appears to be a renewed
interest in the use of entomopathogenic bacteria to control malaria vectors,
which can be attributed to factors related to the fast urbanisation of the
African population and relatively easy access to most breeding sites in urban
areas (Fillinger et al., 2008), to the exophagic behaviour that has been
exhibited by An. gambiae in such areas (Fillinger and Lindsay, 2011;
Geissbuhler et al., 2007; Killeen et al., 2007), and to the insecticide resistance,
mainly to pyrethroids used in insecticide-treated nets, that has emerged for
the primary malaria vectors in many regions of Africa (Cuamba et al.,
2010; Munhenga et al., 2008; Ranson et al., 2011). In view of this scenario,
there is a strong belief that integrated vector management targeting both larval
and adult mosquitoes is the future for malaria control (Townson et al., 2005).
decades ago (Maciel et al., 1996). The Recife City Halls Department of
Health has been running a Filariasis Control Programme, integrating vector
control and mass treatment of humans with diethylcarbamazine since 2003.
Vector-control actions, including Ls treatments, were established in some
critical areas (Cartaxo et al., 2011; Silva-Filha et al., 2008) and were progres-
sively expanded; since 2006, the whole city (94 districts) has been treated.
A conjugate product containing Ls and Bti crystals was also applied in a city
district in 20102011, in an attempt to develop a suitable approach to con-
trol both Cx. quinquefasciatus and Ae. aegypti in urban areas using a single
product that assures effectiveness and low potential for the development
of resistance (C. M. F. Oliveira, personal communication). Another pro-
gramme has been conducted in the Pinheiros River, which crosses the met-
ropolitan area of Sao Paulo (Brazil). This rivers banks are a major source of
Cx. quinquefasciatus in the city, and larvicidal treatment to reduce mosquito
populations is conducted on 22.4 km of the river. From July 2003 to July
2006, 30 treatments were applied, using an Ls-based product in rotation
with a Bti-based larvicide. The larvicides were applied using a boat and
the frequency of treatments is based on surveillance of mosquito immature
stages. No resistance was found in the Cx. quinquefasciatus local population
(Andrade et al., 2007; Silva-Filha et al., 2008). In the United States, Ls and
Bti-based products, as well as conjugate products, have been used in many
counties to fight several species of mosquitoes since the introduction of these
products for vector control.
5. RESISTANCE
5.1. Factors involved in the selection of resistance
The production of biolarvicides based on Ls strains with high larvicidal
action supported their large-scale utilisation in many countries, especially
to fight larvae from the Cx. pipiens complex and anophelines. Field
utilisation showed the effectiveness of this agent but also revealed the onset
of resistance. Some aspects that have contributed to this phenomenon are
mentioned below. The environmental conditions of subtropical and tropical
countries promote mosquito proliferation throughout the year and require
continuous treatment cycles, which increase the selection pressure imposed
on these populations (Barbazan et al., 1997, 1998; Hougard and Back, 1992;
Hougard et al., 1993; Mulla et al., 2001; Regis et al., 1995, 2000b; Silva-
Filha et al., 2001; Skovmand and Bauduin, 1997; Skovmand et al., 2009;
Yuan et al., 2000). The persistence of Ls in some breeding sites (e.g. septic
Microbial Toxins for Mosquito Control 131
tanks) due to its capacity for recycling (Becker et al., 1995; Charles and
Nicolas, 1986; Karch and Coz, 1986) also contributes to a prolonged expo-
sure of larvae to the insecticidal crystals. The adoption of larvicides as the
major tool in control programmes is a strategy that increases the selection
pressure, while the adoption of integrated strategies can significantly reduce
the breeding site area that must be covered by larvicide application. Finally,
the intrinsic mode of action of Ls, based on the action of one major toxin
that targets a single class of receptors (Nielsen-Leroux and Charles, 1992), is
critical for the selection of resistance as reviewed by Wirth (2010) and
Ferreira and Silva-Filha (2013).
five years (Sinegre et al., 1994). This finding, although based on larval sam-
ples from a few breeding sites, revealed for the first time high levels of resis-
tance (RR > 20,000) in field populations. This result raised concerns about
the intensive use of Ls that were confirmed by subsequent resistance reports.
Cx. quinquefasciatus or Cx. pipiens populations from India (Rao et al., 1995),
China (Yuan et al., 2000), Tunisia (Nielsen-Leroux et al., 2002), Thailand
(Mulla et al., 2003), and a second population (BP) from France (Chevillon
et al., 2001; Nielsen-Leroux et al., 2002), exposed to treatments within con-
trol programmes, showed RRs that were, in most cases, higher than 10,000-
fold (Table 3.7). The Ls strains employed in these field-control programmes
were 2362 and 1593, in addition to the C3-41 strain isolated and widely used
for the production of this biolarvicide in China (Yuan et al., 1999). In Brazil,
a Cx. quinquefasciatus population from Coque (Recife) exposed to Ls during
a 2-year period showed a low resistance level (RR 10-fold) (Silva-Filha
et al., 1995). The utilisation of Ls for the control of Cx. quinquefasciatus in
two other urban areas of Agua Fria (Recife) and Rio Pinheiros (Sao Paulo)
in Brazil did not result in the selection of resistance (Silva-Filha et al., 2008).
Factors recorded in those areas, such as the interruption of treatments for
some periods, migration of individuals and rotations with Bti treatment,
might have contributed to the decrease of the selection pressure. Neverthe-
less, the other reports have undoubtedly indicated that high levels of resis-
tance can be achieved in the field, and this resistance remains a central issue
concerning Ls utilisation.
The reports described above indicate that different factors may play an
important role in the development of resistance. The initial susceptibility
of the population to the larvicides used and the frequency of resistance alleles
can be crucial to the evolution of resistance (Andow et al., 2000; Gould et al.,
1997; Tabashnik et al., 2003, 2006). Laboratory Bin-resistant colonies
showed heterogeneous responses when they were subjected to selection
pressure, suggesting that their genetic background could be a determining
factor. Baseline studies of the susceptibility of Cx. pipiens larvae to Ls are
limited; however, those performed with Bti showed variation. The evalua-
tion of approximately 50 Cx. pipiens populations without historic Bti expo-
sure have shown RRs to this agent from less than 3 to 12.5-fold and indicates
pre-existing natural variations (Vasquez et al., 2009; Wirth et al., 2001b).
Generally, the susceptibility of the target populations is not assessed before
the introduction of insecticides in a control programme, although such data
can be critical to detect the development of resistance. For this reason, the
evaluation of larval susceptibility through bioassays and the screening of alleles
associated with Ls resistance in Culex larvae should be routinely performed.
Microbial Toxins for Mosquito Control 135
Cpm1 receptor in Cx. pipiens (Darboux et al., 2001) opened perspectives for
the investigation of the molecular basis of resistance. To date, seven alleles
from cpm1/cqm1 genes associated with Bin resistance were characterised and
the first allele was described in larvae from the GEO colony (Darboux et al.,
2002). The allele cpm1GEO showed a mutation that provoked the loss of the
GPI anchor and prevented the expression of a full Cpm1 midgut-bound
protein. As a consequence, this truncated protein is absent in the epithelial
cell membrane of the midgut and, therefore, cannot act as a receptor for the
Bin toxin. This mechanism consistently explains the high level of resistance
observed for GEO larvae, which is associated with the lack of functional
receptors on the midgut (Nielsen-Leroux et al., 1995; Wirth et al.,
2000b). This pioneer study was followed by the identification of six other
resistance alleles of the cpm1/cqm1 gene in populations from different coun-
tries. Each allele is characterised by a particular mutation and all of them code
for transcripts of truncated or non-functional proteins, which disrupt the Bin
toxin binding to the midgut cells. The mutation in the r alleles associated to
Bin resistance that have been identified to date are shown in Fig. 3.8 and
their features are described below.
The cpm1GEO allele is characterised by a nonsense mutation T1706A that
changes the Leu-569 to a premature stop codon, in the proximity of the puta-
tive GPI signature located in the C-terminus of the Cpm1 protein. This muta-
tion prevents GPI anchoring, while the -glucosidase activity and the toxin
binding capacity of this molecule remains intact, as demonstrated by the
respective recombinant proteins expressed in Sf9 cells (Darboux et al.,
2002). This mutant protein lacks only eleven amino acid residues compared
to the wild-type Cpm1 protein that is composed of 580 amino acids. A second
independent allele was identified in the CqRL1/2362 Cx. quinquefasciatus col-
ony from Brazil that was also characterised by a high level of resistance and lack
of midgut receptors for Bin toxin (Oliveira et al., 2004; Pei et al., 2002). In
larvae from this colony the cqm1REC allele had a deletion of 19 nucleotides (nt)
located from nucleotides 12761294 of its sequence (Romao et al., 2006),
resulting in the reading frame shift that creates a premature stop codon at posi-
tion 1362. The potential truncated protein coded by this transcript would
have only 437 residues and it remains unknown if this protein is produced
in these larvae, since immunodetection assays failed to detect its presence
(Romao et al., 2006). These authors hypothesised that the premature stop
codon could be recognised by the ubiquitous nonsense-mediated decay path-
way of mRNA degradation (Gonzalez et al., 2001; Wagner and Lykke-
Andersen, 2002; Wilkinson and Shyui, 2002) causing the removal of this
transcript.
Microbial Toxins for Mosquito Control 137
Figure 3.8 Nucleotide and deduced amino acid sequence of the open reading frame
from the cqm1 gene, which codes the Cqm1 receptor in Culex quinquefasciatus larvae
(GenBank accession number DQ333335). Seven polymorphisms associated to resistance
to binary toxin that were independently identified in different alleles of the cpm1/cqm1
genes, are represented in this sequence for illustration purpose only. (1) cqm1R allele
shows a deletion of a cytosine (asterisk, boxed) at position 445 that creates a stop codon
downstream (tga, boxed); (2) cpm1BP shows a nonsense mutation that creates a stop
codon (Gln396Stop) at this position (arrow); (3) cpm1BP-del has an alternative splicing
event responsible for the deletion of 66 residues (bold and underlined); (4) cqm1REC
has a 19-nt deletion (bold and underlined); (5) cqm1REC-D25 has a 25-nt deletion
encompassing the 19-nt from the previous deletion (bold and underlined) and six sub-
sequent bases (bold and boxed); (6) cqm1REC-D16 has a 16-nt deletion (bold and italics);
the three last deletions (19-, 25-, 16-nt) create the stop codon located at the same posi-
tion 1362 (tga, boxed); (7) cqm1GEO has a nonsense mutation (T1706A) that creates a
stop codon (Leu569Stop) at this position (arrow).
138 Maria Helena Neves Lobo Silva Filha et al.
phenotype. Nevertheless, in most cases the r alleles have not been identified
and their screening is performed using techniques based on crosses and
followed by the analysis of the progeny susceptibility, such as the F2 and
F1 screens that have been employed to characterise r alleles to Bt toxins
in pest insects (Andow et al., 2000; Gould et al., 1997; Xu et al., 2009;
Yang et al., 2006, 2007a; Zhao et al., 2010). These approaches although effi-
cient, are indirect methods to detect r alleles. On the other hand, the iden-
tification of such alleles allows the development of molecular methods for
their direct detection. Polymerase chain reaction (PCR) methods have been
specifically designed to target specific alleles or to amplify strategic gene seg-
ments that host mutations associated to resistance. Molecular methods for
the detection of mutations in the voltage-gated sodium channel genes that
confer knock-down resistance to pyrethroids have been employed for the
screening of resistance in mosquito populations (Saavedra-Rodriguez
et al., 2007; Sarkar et al., 2009; Singh et al., 2009; Tripet et al., 2006).
For Ls, the identification of alleles of the cpm1/cqm1 gene associated with
resistance opened perspectives for their molecular detection in field
populations. The cqm1REC primarily identified in the CqRL1/2362 Cx.
quinquefasciatus laboratory-selected colony, was detected by allele specific
PCR in all field populations surveyed from Recife city (Brazil), including
six non-treated and one Ls-treated populations. Its frequency in non-treated
samples ranged from 0.001 to 0.017 and the average of five populations was
0.003 (Chalegre et al., 2009, 2012). In three evaluations performed in the
treated area, cqm1REC showed an average frequency of 0.048, which was sig-
nificantly higher than those recorded for the non-treated ones (Chalegre
et al., 2009, 2012). Bioassay analysis performed with these populations
did not reveal RR values significantly between the treated and the non-
treated areas but the early detection of an increased number of individuals
carrying this allele in the treated area is likely to be correlated to the selection
pressure imposed by the Ls exposure. This method can be strategic for the
surveillance of resistance, especially for recessive r alleles. The two other
alleles cqm1REC-D16 and cqm1REC-D25, that were identified in the Cx. quin-
quefasciatus field samples from Recife city, showed a frequency between
0.001 and 0.006 in five populations and a more limited distribution in
the samples compared to the cqm1REC (Chalegre et al., 2012). The study sug-
gests that the cqm1REC is the most important allele associated with resistance
in these populations and it can be used as a marker for the surveillance in that
specific area. To date, the other r alleles to Ls recorded in the literature have
not been tracked in the Cx. pipiens populations.
Microbial Toxins for Mosquito Control 141
A
CqSF CqRL1/2362
L B L B
83
62
48
32
B
CqSF CqRL1/2362
L B L B
83
62
48
32
Guo et al., 2013). The analysis of larvae from the CqRL1/2362 resistant col-
ony showed that the truncated Cqm1 mutant protein does not seem to be
expressed in larvae (Romao et al., 2006). These results suggest that if these
proteins are expressed in larvae it is not certain that they could play their role
in digestion. Biochemical and bioinformatic analyses (Gabrisko, 2013;
Romao et al., 2006) have provided evidence that Cx. quinquefasciatus dis-
plays a set of -glucosidases and the comparative pattern in susceptible
and resistant larvae from the CqRL1/2362 colony are similar, except for
the Cqm1 that is missing in resistant larvae (Fig. 3.9). Maybe the expression
of other -glucosidases in these larvae can compensate the lack of Cqm1 and
Microbial Toxins for Mosquito Control 143
this could be an explanation for the minor level of biological cost observed in
these colonies, although, this hypothesis has not been proved. The specific
role played by Cqm1 and the other -glucosidases in digestion in this species
is still unknown and this would be useful to understand the impact of those
mutations more clearly in the biological fitness of these insects.
The CqRL1/2362 colony, for instance, has been maintained for more than
200 generations under laboratory conditions and the Cqm1 -glucosidase,
which is lacking, does not seem to be essential. A study performed with
the progeny from the cross between resistant CqRL1/2362 and susceptible
individuals in a 1:1 ratio, showed that the resistance alleles displayed a stable
frequency of 0.5 during 10 generations while this colony was kept in the
absence of Ls treatments (Amorim et al., 2010). The biological fitness of
SPHAE and GEO field-derived colonies was not investigated, but these
are also examples of a highly resistant colonies successfully established and
maintained under laboratory conditions during several years (M.H. Silva-
Filha, personal communication). This is consistent with the hypothesis that
some resistance alleles would not be necessarily associated with crucial
adverse effects on biological fitness and they could be maintained in the
populations in the absence of selection pressure (Ffrench-Constant, 2007).
6. MANAGEMENT OF RESISTANCE
6.1. Integrated mosquito-control programmes
The major goal of integrated vector-control approach is to achieve the
reduction of population density, based on the adoption of different methods,
in a cost-effective manner (WHO, 1983, 2004). The utilisation of multiple
strategies to impact the density of mosquitoes, contributes to the decrease of
insecticide utilisation and, thus, for a lower selection pressure. The appear-
ance of resistance has often been associated with the intensive use of larvi-
cides and/or adulticides, as the major intervention adopted to control
insects. This model has predominated since the introduction of synthetic
insecticides and it has greatly contributed to the disseminated resistance
reported worldwide which affects, in particularly, dipteran insects that have
been targeted by control programmes (Brogdon and McAllister, 1998;
Georghiou and Lagunes-Tejeda, 1991; Hemingway et al., 2004; WHO,
1986). In developing countries, the situation is more critical due to lack
of resources and structure that are necessary to set up sustainable
mosquito-control programmes. The choice of suitable insecticides and
144 Maria Helena Neves Lobo Silva Filha et al.
the monitoring of their efficacy are important aspects required for the success
of interventions (Becker et al., 2003). After the large-scale utilisation of Ls
for Cx. quinquefasciatus control, resistance was recorded in exposed
populations in different countries (Section 5), highlighting the need to man-
age the resistance to this control agent.
States, Brazil, India, China and Thailand (Amorim et al., 2007; Nielsen-
Leroux et al., 1995; Rao et al., 1995; Silva-Filha et al., 1995; Su and
Mulla, 2004; Yuan et al., 2003). Individual Cry and Cyt toxins from Bti
or other mosquitocidal Bt strains, such as serovar. medellin (Btmed) and
serovar. jegathesan (Btjeg) are able to overcome Bin toxin resistance. The syn-
ergistic interactions among these toxins and Bin toxin have been demon-
strated using mixtures of toxins produced in the respective native strains,
or by their combined expression in recombinant Bti or Ls strains as
described below.
The synergistic effect of Cyt toxins on Bin toxicity was first investigated
using the Cyt1Ab1 from Btmed. In this case, a recombinant Ls 2297 strain
that produced both Bin and Cyt1Ab1 exhibited a significant activity against
the larvae from GEO- and SPHAE-resistant colonies (Thiery et al., 1998).
Mixtures of Cyt1Aa or Cyt2Ba, from Bti, with the Bin toxin were shown to
be effective against Ae. aegypti larvae that is a refractory species to Bin toxin
(Wirth et al., 2000a) and to Cx. pipiens larvae that were resistant to the Bin
toxin (Wirth et al., 2000c, 2001a, 2004). A laboratory selection trial of Cx.
quinquefasciatus larvae using a mixture of Bin and Cyt1 toxins in a 3:1 ratio
during 20 generations showed a RR of 1.4 while a RR > 1000 was observed
for their counterparts selected with the Bin toxin only (Wirth et al., 2005),
corroborating the role of the Cyt1Aa toxin to delay Bin toxin resistance.
The synergy of the whole Bti crystal, or its individual Cry toxins, with
Bin toxin has also been demonstrated. The introduction of Bti toxins in Ls
strains has been attempted, before the advent of Bin toxin resistance, aiming
at the improving the spectrum of Ls toxicity since this is more restricted than
Bti (Bar et al., 1991; Trisrisook et al., 1990). Later, working with Bin resis-
tance, a study evaluated a set of mixtures of Bin with Bti and its toxins that
showed an enhanced activity to Bin-resistant larvae and the potential of
these mixtures to overcome the resistance to the Bin toxin (Wirth et al.,
2004). Evaluation of the synergism of six different mixtures of wild Ls
and Bti crystals to susceptible Cx. quinquefasciatus showed that the highest syn-
ergism factor found was observed using a 3:1 ratio of Ls to Bti (Sreshty et al.,
2011). This combination also displayed faster cytopathological effects on the
midgut epithelium and muscles.
The production of recombinant Ls strains containing Bti toxins and/or
the Cry11Ba toxin from Btjeg showed an enhanced activity to larvae from
Bin-resistant colonies and partially restored their susceptibility to the Bin
toxin (Poncet et al., 1994, 1997; Servant et al., 1999). However, it is impor-
tant to note that the introduction of Bt genes in Ls strains has been often
Microbial Toxins for Mosquito Control 147
Environmental management
Use of L. sphaericus within
integrated programme Other mosquito-control agents
Personnal protection
Figure 3.10 Strategies for preventing the resistance to the Bin toxin from Lysinibacillus
sphaericus (Ls) and agents/toxins that can be used for its management. Insect growth
regulator (IGR); Bacillus thuringiensis serovar. israelensis (Bti), serovar. jegathesan (Btjeg),
serovar. medellin (Btmed); mosquitocidal toxins (Mtxs).
150 Maria Helena Neves Lobo Silva Filha et al.
(Cetin et al., 2005; Giraldo-Calderon et al., 2008; Guidi et al., 2013b; Liu
et al., 2004; Marcombe et al., 2011; Pridgeon et al., 2008; Tetreau et al.,
2013). Spinosins, for instance, are also bacterial larvicides that have been
recently registered in some countries for mosquito control (Hertlein
et al., 2010). They show a relatively narrow toxicity spectrum, and have
recently been introduced with promising results, alone or combination with
other biolarvicides (Anderson et al., 2011; Cetin et al., 2005; Harbison et al.,
2013; Jiang and Mulla, 2009; Kumar et al., 2011; Marcombe et al., 2011;
Marina et al., 2012). The rational utilisation of Ls in the context of an inte-
grated programme using suitable control strategies and including monitoring
of the mosquito susceptibility can ensure its effectiveness for the control
interventions and can prevent the onset of resistance. The major aspects
concerning the management of resistance to Ls presented in this section
are summarised in Fig. 3.10.
ACKNOWLEDGEMENTS
We thank all colleagues who provided suggestions for this work; publishers for the
permissions for using figures presented in this chapter; Claudia Maria Fontes de Oliveira,
Const^ancia Junqueira Ayres, Maria Alice Varjal de Melo-Santos and Ros^angela Maria
Barbosa from the Department of Entomology (CPqAM-FIOCRUZ) for the
encouragement for writing this chapter.
REFERENCES
Abdullah, M.A., Valaitis, A.P., Dean, D.H., 2006. Identification of a Bacillus thuringiensis
Cry11Ba toxin-binding aminopeptidase from the mosquito Anopheles quadrimaculatus.
BMC Biochem. 7, 16.
Ahmed, H.K., Mitchell, W.J., Priest, F.G., 1995. Regulation of mosquitocidal toxin synthe-
sis in Bacillus sphaericus. Appl. Microbiol. Biotechnol. 43, 310314.
Ahmed, I., Yokota, A., Yamazoe, A., Fujiwara, T., 2007. Proposal of Lysinibacillus
boronitolerans gen. nov. sp. nov., and transfer of Bacillus fusiformis to Lysinibacillus fusiformis
comb. nov. and Bacillus sphaericus to Lysinibacillus sphaericus comb. nov. Int. J. Syst.
Evol. Microbiol. 57, 11171125.
Alexander, B., Priest, F.G., 1990. Numerical classification and identification of Bacillus
sphaericus including some strains pathogenic for mosquito larvae. J. Gen. Microbiol.
136, 367376.
Almog, O., Gonzalez, A., Klein, D., Greenblatt, H.M., Braun, S., Shoham, G., 2003. The
0.93A crystal structure of sphericase: a calcium-loaded serine protease from Bacillus
sphaericus. J. Mol. Biol. 332, 10711082.
Aly, C., Mulla, M.S., Federici, B.A., 1989. Ingestion, dissolution, and proteolysis of the Bacil-
lus sphaericus toxin by mosquito larvae. J. Invertebr. Pathol. 53, 1220.
Amorim, L.B., Oliveira, C.M.F., Rios, E.M., Regis, L., Silva-Filha, M.H.N.L., 2007.
Development of Culex quinquefasciatus resistance to Bacillus sphaericus strain IAB59 needs
long term selection pressure. Biol. Control 42, 155160.
Microbial Toxins for Mosquito Control 151
Amorim, L.B., De Barros, R.A., Chalegre, K.D., De Oliveira, C.M., Regis, L.N., Silva-
Filha, M.H., 2010. Stability of Culex quinquefasciatus resistance to Bacillus sphaericus eval-
uated by molecular tools. Insect Biochem. Mol. Biol. 40, 311316.
Anderson, J.F., Ferrandino, F.J., Dingman, D.W., Main, A.J., Andreadis, T.G., Becnel, J.J.,
2011. Control of mosquitoes in catch basins in Connecticut with Bacillus thuringiensis
israelensis, Bacillus sphaericus, [corrected] and spinosad. J. Am. Mosq. Control Assoc.
27, 4555.
Andow, D.A., Olson, D.M., Hellmich, R.L., Alstad, D.N., Hutchison, W.D., 2000.
Frequency of resistance to Bacillus thuringiensis toxin Cry1Ab in an Iowa population of
European corn borer (Lepidoptera: Crambidae). J. Econ. Entomol. 93, 2630.
Andrade, C.F.S., Campos, J.C., Cabrini, I., Marques Filho, C.A.M., Hibi, S., 2007.
Suscetibilidade de populacoes de Culex quinquefasciatus Say (Diptera: Culicidae) sujeitas
ao controle com Bacillus sphaericus Neide no rio Pinheiros, Sao Paulo. BioAssay 2, 14.
Anilkumar, K.J., Pusztai-Carey, M., Moar, W.J., 2008. Fitness costs associated with Cry1Ac-
resistant Helicoverpa zea (Lepidoptera: Noctuidae): a factor countering selection for resis-
tance to Bt cotton? J. Econ. Entomol. 101, 14211431.
Aquino De Muro, M., Mitchell, W.J., Priest, F.G., 1992. Differentiation of mosquito-
pathogenic strains of Bacillus sphaericus from non-toxic varieties by ribosomal RNA gene
restriction patterns. J. Gen. Microbiol. 138, 11591166.
Araujo, A.P., Melo-Santos, M.A.V., Carlos, S.O., Rios, E.M., Regis, L., 2007. Evaluation of
an experimental product based on Bacillus thuringiensis sorovar. israelensis against Aedes
aegypti larvae (Diptera: Culicidae). Biol. Control 41, 339347.
Arenas, I., Bravo, A., Soberon, M., Gomez, I., 2010. Role of alkaline phosphatase from Man-
duca sexta in the mechanism of action of Bacillus thuringiensis Cry1Ab toxin. J. Biol.
Chem. 285, 1249712503.
Arredondo-Jimenez, J.I., Lopez, T., Rodriguez, M.H., Bown, D.N., 1990. Small scale field
trials of Bacillus sphaericus (strain 2362) against anopheline and culicine mosquito larvae in
southern Mexico. J. Am. Mosq. Control Assoc. 6, 300305.
Atsumi, S., Miyamoto, K., Yamamoto, K., Narukawa, J., Kawai, S., Sezutsu, H.,
Kobayashi, I., Uchino, K., Tamura, T., Mita, K., Kadono-Okuda, K., Wada, S.,
Kanda, K., Goldsmith, M.R., Noda, H., 2012. Single amino acid mutation in an
ATP-binding cassette transporter gene causes resistance to Bt toxin Cry1Ab in the silk-
worm Bombyx mori. Proc. Natl. Acad. Sci. U.S.A 109, E1591E1598.
Bar, E., Lieman-Hurwitz, J., Rahamim, E., Keynan, A., Sandler, N., 1991. Cloning and
expression of Bacillus thuringiensis israelensis delta-endotoxin DNA in B. sphaericus.
J. Invertebr. Pathol. 57, 149158.
Barbazan, P., Baldet, T., Darriet, F., Escaffre, H., Djoda, D.H., Hougard, J.M., 1997.
Control of Culex quinquefasciatus (Diptera: Culicidae) with Bacillus sphaericus in Maroua,
Cameroon. J. Am. Mosq. Control Assoc. 13, 263269.
Barbazan, P., Baldet, T., Darriet, F., Escaffre, H., Djoda, D.H., Hougard, J.M., 1998. Impact
of treatments with Bacillus sphaericus on Anopheles populations and the transmission of
malaria in Maroua, a large city in a savannah region of Cameroon. J. Am. Mosq. Control
Assoc. 14, 3339.
Baumann, L., Baumann, P., 1991. Effects of components of the Bacillus sphaericus toxin on
mosquito larvae and mosquito-derived tissue culture-grown cells. Curr. Microbiol.
23, 5157.
Baumann, P., Unterman, B.M., Baumann, L., Broadwell, A.H., Abbene, S.J.,
Bowditch, R.D., 1985. Purification of the larvicidal toxin of Bacillus sphaericus and
evidence for high-molecular-weight precursors. J. Bacteriol. 163, 738747.
Baumann, P., Baumann, L., Bowditch, R.D., Broadwell, A.H., 1987. Cloning of the gene
for the larvicidal toxin of Bacillus sphaericus 2362: evidence for a family of related
sequences. J. Bacteriol. 169, 40614067.
152 Maria Helena Neves Lobo Silva Filha et al.
Baumann, L., Broadwell, A.H., Baumann, P., 1988. Sequence analysis of the mosquitocidal
toxin genes encoding 51.4- and 41.9-kilodalton proteins from Bacillus sphaericus 2362 and
2297. J. Bacteriol. 170, 20452050.
Baumann, P., Clark, M.A., Baumann, L., Broadwell, A.H., 1991. Bacillus sphaericus as a mos-
quito pathogen: properties of the organism and its toxins. Microbiol. Rev. 55, 425436.
Becker, N., 1997. Microbial control of mosquitoes: management of the upper Rhine mos-
quito population as a model programme. Parasitol. Today 13, 485487.
Becker, N., 1998. The use of Bacillus thuringiensis subs. israelensis (Bti) against mosquitoes,
with special emphasis on the ecological impact. Isr. J. Entomol. 32, 6369.
Becker, N., 2000. Bacterial control of vector mosquitoes and black flies. In: Charles, J.F.,
Delecluse, A., Nielsen-Leroux, C. (Eds.), Entomopathogenic Bacteria: From Laboratory
to Filed Application. Kluwer Academic Publishers, Dordrecht, pp. 384398.
Becker, N., 2003. Ice granules containing endotoxins of microbial agents for the control of
mosquito larvaea new application technique. J. Am. Mosq. Control Assoc. 19, 6366.
Becker, N., Zgomba, M., Petric, D., Beck, M., Ludwig, M., 1995. Role of larval cadavers in
recycling processes of Bacillus sphaericus. J. Am. Mosq. Control Assoc. 11, 329334.
Becker, N., Petric, D., Dahl, C., Lane, J., Kaiser, A., 2003. Integrated pest management.
In: Becker, N. (Ed.), Mosquitos and Their Control. Kluwer Academic/Plenum
Publishers, New York, pp. 417424.
Berrocal, E., Carey, C., Rodriguez, L., Calampa, C., Valdivia, L., 2000. A residual effect of
Bacillus sphaericus for control of Anopheles darlingi, in Iquitos in Peru. J. Am. Mosq. Con-
trol Assoc. 16, 295312.
Berry, C., 2012. The bacterium, Lysinibacillus sphaericus, as an insect pathogen. J. Invertebr.
Pathol. 109, 110.
Berry, C., Hindley, J., 1987. Bacillus sphaericus strain 2362: identification and nucleotide
sequence of the 41.9 kDa toxin gene. Nucleic Acids Res. 15, 5891.
Berry, W.J., Novak, M.G., Khounlo, S., Rowley, W.A., Melchior, G.L., 1987. Efficacy of
Bacillus sphaericus and Bacillus thuringiensis var. israelensis for control of Culex pipiens and
floodwater Aedes larvae in Iowa. J. Am. Mosq. Control Assoc. 3, 579582.
Berry, C., Hindley, J., Ehrhardt, A.F., Grounds, T., De Souza, I., Davidson, E.W., 1993.
Genetic determinants of host ranges of Bacillus sphaericus mosquito larvicidal toxins.
J. Bacteriol. 175, 510518.
Berry, C., Oneil, S., Ben-Dov, E., Jones, A.F., Murphy, L., Quail, M.A., Holden, M.T.,
Harris, D., Zaritsky, A., Parkhill, J., 2002. Complete sequence and organization of
pBtoxis, the toxin-coding plasmid of Bacillus thuringiensis subsp. israelensis. App. Environ.
Microbiol. 68, 50825095.
Berti, J., Gonzales, J., 2004. Evaluacion de la efectividad y persistencia de la nueva
formulacion de Bacillus sphaericus contra larvas de Anopheles aquasalis Curry en los
criaderos naturales del estado Sucre, Venezuela. Bol. Mal. Salud Amb. 44, 2127.
Blanco-Castro, S.D., Martnez Arias, A., Cano Velasquez, O.R., Tello Granados, R.,
Mendonza, I., 2000. Introduction of Bacillus sphaericus strain-2362 (GRISELESF) for
biological control of malaria vectors in Guatemala. Rev. Cubana Med. Trop. 52, 3743.
Bone, L.W., Tinelli, R., 1987. Trichostrongylus colubriformis: larvicidal activity of toxic extracts
from Bacillus sphaericus (strain 1593) spores. Exp. Parasitol. 64, 514516.
Boonserm, P., Moonsom, S., Boonchoy, C., Promdonkoy, B., Parthasarathy, K., Torres, J.,
2006. Association of the components of the binary toxin from Bacillus sphaericus in
solution and with model lipid bilayers. Biochem. Biophys. Res. Commun. 342,
12731278.
Boonyos, P., Soonsanga, S., Boonserm, P., Promdonkoy, B., 2010. Role of cysteine at posi-
tions 67, 161 and 241 of a Bacillus sphaericus binary toxin BinB. BMB Rep. 43, 2328.
Bourdeau, R.W., Malito, E., Chenal, A., Bishop, B.L., Musch, M.W., Villereal, M.L.,
Chang, E.B., Mosser, E.M., Rest, R.F., Tang, W.J., 2009. Cellular functions and
Microbial Toxins for Mosquito Control 153
Cetin, H., Yanikoglu, A., Cilek, J.E., 2005. Evaluation of the naturally-derived insecticide
spinosad against Culex pipiens L. (Diptera: Culicidae) larvae in septic tank water in
Antalya, Turkey. J. Vector Ecol. 30, 151154.
Cetin, H., Dechant, P., Yanikoglu, A., 2007. Field trials with tank mixtures of Bacillus thur-
ingiensis subsp. israelensis and Bacillus sphaericus formulations against Culex pipiens larvae in
septic tanks in Antalya, Turkey. J. Am. Mosq. Control Assoc. 23, 161165.
Chalegre, K.D., Romao, T.P., Amorim, L.B., Anastacio, D.B., De Barros, R.A., De
Oliveira, C.M., Regis, L., De-Melo-Neto, O.P., Silva-Filha, M.H., 2009. Detection
of an allele conferring resistance to Bacillus sphaericus binary toxin in Culex quin-
quefasciatus populations by molecular screening. Appl. Environ. Microbiol. 75,
10441049.
Chalegre, K.D., Romao, T.P., Tavares, D.A., Santos, E.M., Ferreira, L.M., Oliveira, C.M.
F., De-Melo-Neto, O.P., Silva-Filha, M.H.N.L., 2012. Novel mutations associated to
Bacillus sphaericus resistance are identified in a polymorphic region of the Culex quin-
quefasciatus cqm1 gene Appl. Environ. Microbiol. 78, 63216326.
Chan, S.W., Thanabalu, T., Wee, B.Y., Porter, A.G., 1996. Unusual amino acid determi-
nants of host range in the Mtx2 family of mosquitocidal toxins. J. Biol. Chem.
271, 1418314187.
Charles, J.F., 1987. Ultrastructural midgut events in Culicidae larvae fed with Bacillus
sphaericus 2297 spore/crystal complex. Ann. Inst. Pasteur Microbiol. 138, 471484.
Charles, J.F., Nicolas, L., 1986. Recycling of Bacillus sphaericus 2362 in mosquito larvae:
a laboratory study. Ann. Inst. Pasteur Microbiol. 137B, 101111.
Charles, J.F., Nielsen-Leroux, C., 2000. Mosquitocidal bacterial toxins: diversity, mode of
action and resistance phenomena. Mem. Inst. Oswaldo Cruz 95 (Suppl. 1), 201206.
Charles, J.F., Kalfon, A., Bourgouin, C., de Barjac, H., 1988. Bacillus sphaericus asporogenous
mutants: morphology, protein pattern and larvicidal activity. Ann. Inst. Pasteur
Microbiol. 139, 243259.
Charles, J.F., Hamon, S., Baumann, P., 1993. Inclusion bodies and crystals of Bacillus
sphaericus mosquitocidal proteins expressed in various bacterial hosts. Res. Microbiol.
144, 411416.
Charles, J.F., Nielsen-Leroux, C., Delecluse, A., 1996. Bacillus sphaericus toxins: molecular
biology and mode of action. Annu. Rev. Entomol. 41, 451472.
Charles, J.F., Silva-Filha, M.H., Nielsen-Leroux, C., Humphreys, M.J., Berry, C., 1997.
Binding of the 51- and 42-kDa individual components from the Bacillus sphaericus crystal
toxin to mosquito larval midgut membranes from Culex and Anopheles sp. (Diptera:
Culicidae). FEMS Microbiol. Lett. 156, 153159.
Charles, J.F., Darboux, I., Pauron, D., Nielsen-Leroux, C., 2010. Mosquitocidal Bacillus
sphaericus: toxins, genetics, mode of action, use, and resistance mechanisms.
In: Gilbert, L., Gill, S.S. (Eds.), Insect Control: Biological and Synthetic Agents.
Elsevier, London, pp. 283307.
Chen, J., Aimanova, K.G., Fernandez, L.E., Bravo, A., Soberon, M., Gill, S.S., 2009a. Aedes
aegypti cadherin serves as a putative receptor of the Cry11Aa toxin from Bacillus thur-
ingiensis subsp. israelensis. Biochem. J. 424, 191200.
Chen, J., Aimanova, K.G., Pan, S., Gill, S.S., 2009b. Identification and characterization of
Aedes aegypti aminopeptidase N as a putative receptor of Bacillus thuringiensis Cry11A
toxin. Insect Biochem. Mol. Biol. 39, 688696.
Cheong, W.C., Yap, H.H., 1985. Bioassays of Bacillus sphaericus (strain 1593) against mos-
quitoes of public health importance in Malaysia. Southeast Asian J. Trop. Med. Public
Health 16, 5458.
Chevillon, C., Bernard, C., Marquine, M., Pasteur, N., 2001. Resistance to Bacillus sphaericus
in Culex pipiens (Diptera: Culicidae): interaction between recessive mutants and evolu-
tion in southern France. J. Med. Entomol. 38, 657664.
Microbial Toxins for Mosquito Control 155
Chiou, C., Davidson, E.W., Thanabalu, T., Porter, A.G., Allen, J.P., 1999. Crystallization and
preliminary X-ray diffraction studies of the 51 kDa protein of the mosquito-larvicidal
binary toxin from Bacillus sphaericus. Acta Crystallogr. D Biol. Crystallogr. 55, 10831085.
Clark, M.A., Baumann, P., 1990. Deletion analysis of the 51-kilodalton protein of the Bacillus
sphaericus 2362 binary mosquitocidal toxin: construction of derivatives equivalent to the
larva-processed toxin. J. Bacteriol. 172, 67596763.
Clark, M.A., Baumann, P., 1991. Modification of the Bacillus sphaericus 51- and 42-kilodalton
mosquitocidal proteins: effects of internal deletions, duplications, and formation of
hybrid proteins. Appl. Environ. Microbiol. 57, 267271.
Cokmus, C., Davidson, E.W., Cooper, K., 1997. Electrophysiological effects of Bacillus
sphaericus binary toxin on cultured mosquito cells. J. Invertebr. Pathol. 69, 197204.
Consoli, R.A., Santos Bde, S., Lamounier, M.A., Secundino, N.F., Rabinovitch, L.,
Silva, C.M., Alves, R.S., Carneiro, N.F., 1997. Efficacy of a new formulation of Bacillus
sphaericus 2362 against Culex quinquefasciatus (Diptera: Culicidae) in Montes Claros,
Minas Gerais, Brazil. Mem. Inst. Oswaldo Cruz 92, 571573.
Correa, M., Yousten, A.A., 1995. Bacillus sphaericus spore germination and recycling in mos-
quito larval cadavers. J. Invertebr. Pathol. 66, 7681.
Cuamba, N., Morgan, J.C., Irving, H., Steven, A., Wondji, C.S., 2010. High level of pyre-
throid resistance in an Anopheles funestus population of the Chokwe district in Mozam-
bique. PLoS One 5, e11010.
Darboux, I., Nielsen-Leroux, C., Charles, J.F., Pauron, D., 2001. The receptor of Bacillus
sphaericus binary toxin in Culex pipiens (Diptera: Culicidae) midgut: molecular cloning
and expression. Insect Biochem. Mol. Biol. 31, 981990.
Darboux, I., Pauchet, Y., Castella, C., Silva-Filha, M.H., Nielsen-Leroux, C., Charles, J.F.,
Pauron, D., 2002. Loss of the membrane anchor of the target receptor is a mechanism of
bioinsecticide resistance. Proc. Natl. Acad. Sci. U.S.A 99, 58305835.
Darboux, I., Charles, J.F., Pauchet, Y., Warot, S., Pauron, D., 2007. Transposon-mediated
resistance to Bacillus sphaericus in a field-evolved population of Culex pipiens (Diptera:
Culicidae). Cell. Microbiol. 9, 20222029.
Davidson, E.W., 1979. Ultrastructure of midgut events in the pathogenesis of Bacillus
sphaericus strain SSII-1 infections of Culex pipiens quinquefasciatus larvae. Can. J.
Microbiol. 25, 178184.
Davidson, E.W., 1988. Binding of the Bacillus sphaericus (Eubacteriales: Bacillaceae) toxin to
midgut cells of mosquito (Diptera: Culicidae) larvae: relationship to host range. J. Med.
Entomol. 25, 151157.
Davidson, E.W., 1989. Variation in binding of Bacillus sphaericus toxin and wheat germ agglu-
tinin to larval midgut cells of six species of mosquitoes. J. Invertebr. Pathol. 53, 251259.
Davidson, E.W., Myers, P., 1981. Parasporal inclusions in Bacillus sphaericus. FEMS
Microbiol. Lett. 10, 261265.
Davidson, E.W., Titus, M., 1987. Ultrastructural effects of the Bacillus sphaericus mosquito
larvicidal toxin on cultured mosquito cells. J. Invertebr. Pathol. 50, 213220.
Davidson, E.W., Morthon, H.L., Moffett, J.O., Singer, S., 1977. Effect of Bacillus sphaericus
strain SSII-1 on honey bees, Apis mellifera. J. Invertebr. Pathol. 29, 344346.
Davidson, E.W., Urbina, M., Payne, J., Mulla, M.S., Darwazeh, H., Dulmage, H.T.,
Correa, J.A., 1984. Fate of Bacillus sphaericus 1593 and 2362 spores used as larvicides
in the aquatic environment. Appl. Environ. Microbiol. 47, 125129.
Davidson, E.W., Shellabarger, C., Meyer, M., Bieber, A.L., 1987. Binding of the Bacillus
sphaericus mosquito larvicidal toxin to cultured insect cells. Can. J. Microbiol. 33, 982989.
Davidson, E.W., Oei, C., Meyer, M., Bieber, A.L., Hindley, J., Berry, C., 1990. Interaction
of the Bacillus sphaericus mosquito larvicidal proteins. Can. J. Microbiol. 36, 870878.
De Barjac, H., 1978. Toxicity of Bacillus thuringiensis var. israelensis for larvae of Aedes aegypti and
Anopheles stephensi. C. R. Acad. Sci. Hebd. Seances Acad. Sci. D 286, 11751178.
156 Maria Helena Neves Lobo Silva Filha et al.
De Barjac, H., 1990. Classification of Bacillus sphaericus trains and comparative toxicity
to mosquito larvae. In: De Barjac, H., Sutherland, D. (Eds.), Bacterial Control of
Mosquitoes and Black-Flies: Biochemistry, Genetics, and Applications of Bacillus
thuringiensis and Bacillus sphaericus. Rutgers University Press, New Brunswick,
pp. 228236.
De Barjac, H., Charles, J.F., 1983. Une nouvelle toxine active sur les moustiques, presente
dans des inclusions crystallines produites para Bacillus sphaericus. C. R. Acad. Sci. Paris Ser.
III 296, 905910.
De Barjac, H., Veron, M.C., Dumanoir, V., 1980. Biochemical and serological characteri-
zation of Bacillus sphaericus strains, pathogenic or non-pathogenic for mosquitoes
(authors transl). Ann. Microbiol. 131B, 191201.
De Barjac, H., Larget-Thiery, I., Cosmao Dumanoir, V., Ripouteau, H., 1985. Serological
classification of Bacillus sphaericus strains in relation with toxicity to mosquito larvae. Appl.
Microbiol. Biotechnol. 21, 8590.
De La Torre, F., Bennardo, T., Sebo, P., Szulmajster, J., 1989. On the respective roles of the
two proteins encoded by the Bacillus sphaericus 1593M toxin genes expressed in Escherichia
coli and Bacillus subtilis. Biochem. Biophys. Res. Commun. 164, 14171422.
De Maagd, R.A., Bravo, A., Berry, C., Crickmore, N., Schnepf, H.E., 2003. Structure,
diversity, and evolution of protein toxins from spore-forming entomopathogenic bacte-
ria. Annu. Rev. Genet. 37, 409433.
De Melo, J.V., Vasconcelos, R.H., Furtado, A.F., Peixoto, C.A., Silva-Filha, M.H., 2008.
Ultrastructural analysis of midgut cells from Culex quinquefasciatus (Diptera: Culicidae)
larvae resistant to Bacillus sphaericus. Micron 39, 13421350.
De Melo, J.V., Jones, G.W., Berry, C., Vasconcelos, R.H., De Oliveira, C.M., Furtado, A.F.,
Peixoto, C.A., Silva-Filha, M.H., 2009. Cytopathological effects of Bacillus sphaericus
Cry48Aa/Cry49Aa toxin on binary toxin-susceptible and -resistant Culex quinquefasciatus
larvae. Appl. Environ. Microbiol. 75, 47824789.
De Oliveira, C.M., Filho, F.C., Beltran, J.E., Silva-Filha, M.H., Regis, L., 2003. Biological
fitness of a Culex quinquefasciatus population and its resistance to Bacillus sphaericus. J. Am.
Mosq. Control Assoc. 19, 125129.
Delecluse, A., Juarez-Perez, V., Berry, C., 2000. Vector-active toxins: structure and diver-
sity. In: Charles, J.F., Delecluse, A., Nielsen-Leroux, C. (Eds.), Entomopathogenic Bac-
teria: From Laboratory to Field Application. Kluwer Academic Publishers, Dordrecht,
pp. 101125.
Dennett, J.A., Meisch, M.V., 2000. Effectiveness of aerial- and ground-applied Bacillus for-
mulations against Anopheles quadrimaculatus larvae in Arkansas rice plots. J. Am. Mosq.
Control Assoc. 16, 229233.
Dennett, J.A., Meek, C.L., Meisch, M.V., 2001. Efficacy of VectoLex WDG against Anoph-
eles quadrimaculatus and Psorophora columbiae larvae in Arkansas and Mississippi rice. J. Am.
Mosq. Control Assoc. 17, 231237.
Des Rochers, B., Garcia, R., 1984. Evidence for persistence and recycling of Bacillus
sphaericus. Mosq. News 44, 160165.
Dritz, D.A., Lawler, S.P., Evkhanian, C., Graham, P., Baracosa, V., Dula, G., 2011. Control
of mosquito larvae in seasonal wetlands on a wildlife refuge using VectoMax CG. J. Am.
Mosq. Control Assoc. 27, 398403.
Du, Y., Nomura, Y., Satar, G., Hu, Z., Nauen, R., He, S.Y., Zhorov, B.S., Dong, K., 2013.
Molecular evidence for dual pyrethroid-receptor sites on a mosquito sodium channel.
Proc. Natl. Acad. Sci. U.S.A 110, 1178511790.
Elangovan, G., Shanmugavelu, M., Rajamohan, F., Dean, D.H., Jayaraman, K., 2000.
Identification of the functional site in the mosquito larvicidal binary toxin of Bacillus
sphaericus 1593M by site-directed mutagenesis. Biochem. Biophys. Res. Commun.
276, 10481055.
Microbial Toxins for Mosquito Control 157
El-Bendary, M., Priest, F.G., Charles, J.F., Mitchell, W.J., 2005. Crystal protein synthesis is
dependent on early sporulation gene expression in Bacillus sphaericus. FEMS Microbiol.
Lett. 252, 5156.
Eritja, R., 2013. Laboratory tests on the efficacy of VBC60035, a combined larvicidal for-
mulation of Bacillus thuringiensis israelensis (strain AM65-52) and Bacillus sphaericus (strain
2362) against Aedes albopictus in simulated catch basins. J. Am. Mosq. Control Assoc.
29, 280283.
Federici, B.A., Park, H.W., Bideshi, D.K., Wirth, M.C., Johnson, J.J., 2003. Recombinant
bacteria for mosquito control. J. Exp. Biol. 206, 38773885.
Federici, B.A., Park, H.W., Bideshi, D.K., Wirth, M.C., Johnson, J.J., Sakano, Y., Tang, M.,
2007. Developing recombinant bacteria for control of mosquito larvae. J. Am. Mosq.
Control Assoc. 23, 164175.
Federici, B.A., Park, H.D., Bideshi, D.K., 2010. Overview of the basic biology of Bacillus
thuringiensis with emphasis on genetic engineering of bacterial larvicides for mosquito
control. Open J. Toxicol. 3, 83100.
Fernandez, L.E., Aimanova, K.G., Gill, S.S., Bravo, A., Soberon, M., 2006. A GPI-anchored
alkaline phosphatase is a functional midgut receptor of Cry11Aa toxin in Aedes aegypti
larvae. Biochem. J. 394, 7784.
Fernandez-Luna, M.T., Lanz-Mendoza, H., Gill, S.S., Bravo, A., Soberon, M., Miranda-
Rios, J., 2010. An alpha-amylase is a novel receptor for Bacillus thuringiensis ssp. israelensis
Cry4Ba and Cry11Aa toxins in the malaria vector mosquito Anopheles albimanus (Diptera:
Culicidae). Environ. Microbiol. 12, 746757.
Ferreira, L.M., Silva-Filha, M.H.N.L., 2013. Bacterial larvicides for vector control: mode of
action of toxins and implications for resistance. Biocontrol Sci. Technol. 23, 11371168.
Ferreira, L.M., Romao, T.P., De-Melo-Neto, O.P., Silva-Filha, M.H., 2010. The
orthologue to the Cpm1/Cqm1 receptor in Aedes aegypti is expressed as a midgut
GPI-anchored alpha-glucosidase, which does not bind to the insecticidal binary toxin.
Insect Biochem. Mol. Biol. 40, 604610.
Ferreira, L.M., Romao, T.P., Nascimento, N.A., Costa, M.D., Rezende, A.M., De-Melo-
Neto, O.P., Silva-Filha, M.H., 2014. Non conserved residues between Cqm1 and Aam1
mosquito alpha-glucosidases are critical for the capacity of Cqm1 to bind the Binary (Bin)
toxin from Lysinibacillus sphaericus. Insect Biochem. Mol. Biol. 50, 3442.
Ffrench-Constant, R.H., 2007. Which came first: insecticides or resistance? Trends Genet.
23, 14.
Fillinger, U., Lindsay, S.W., 2006. Suppression of exposure to malaria vectors by an order of
magnitude using microbial larvicides in rural Kenya. Trop. Med. Int. Health
11, 16291642.
Fillinger, U., Lindsay, S.W., 2011. Larval source management for malaria control in Africa:
myths and reality. Malar. J. 10, 353.
Fillinger, U., Knols, B.G., Becker, N., 2003. Efficacy and efficiency of new Bacillus thur-
ingiensis var israelensis and Bacillus sphaericus formulations against Afrotropical anophelines
in Western Kenya. Trop. Med. Int. Health 8, 3747.
Fillinger, U., Kannady, K., William, G., Vanek, M.J., Dongus, S., Nyika, D.,
Geissbuhler, Y., Chaki, P.P., Govella, N.J., Mathenge, E.M., Singer, B.H.,
Mshinda, H., Lindsay, S.W., Tanner, M., Mtasiwa, D., et al., 2008. A tool box for oper-
ational mosquito larval control: preliminary results and early lessons from the Urban
Malaria Control Programme in Dar es Salaam, Tanzania. Malar. J. 7, 20.
Fillinger, U., Ndenga, B., Githeko, A., Lindsay, S.W., 2009. Integrated malaria vector con-
trol with microbial larvicides and insecticide-treated nets in western Kenya: a controlled
trial. Bull. WHO 87, 655665.
Floore, T., Wardz, R., 2009. Evaluation of Bacillus sphaericus against Mansonia dyari larvae in
phosphate lakes in Polk County. Florida. J. Am. Mosq. Control Assoc. 25, 310314.
158 Maria Helena Neves Lobo Silva Filha et al.
Frachon, E., Hamon, S., Nicolas, L., de Barjac, H., 1991. Cellular fatty acid analysis as a
potential tool for predicting mosquitocidal activity of Bacillus sphaericus strains. Appl.
Environ. Microbiol. 57, 33943398.
From, C., Granum, P.E., Hardy, S.P., 2008. Demonstration of a cholesterol-dependent
cytolysin in a noninsecticidal Bacillus sphaericus strain and evidence for widespread distri-
bution of the toxin within the species. FEMS Microbiol. Lett. 286, 8592.
Gabrisko, M., 2013. Evolutionary history of eukaryotic alpha-glucosidases from the alpha-
amylase family. J. Mol. Evol. 76, 129145.
Gahan, L.J., Pauchet, Y., Vogel, H., Heckel, D.G., 2010. An ABC transporter mutation is
correlated with insect resistance to Bacillus thuringiensis Cry1Ac toxin. PLoS Genet.
6, e1001248.
Galardo, A.K., Zimmerman, R., Galardo, C.D., 2013. Larval control of Anopheles
(Nyssorhinchus) darlingi using granular formulation of Bacillus sphaericus in abandoned
gold-miners excavation pools in the Brazilian Amazon rainforest. Rev. Soc. Bras.
Med. Trop. 46, 172177.
Gammon, K., Jones, G.W., Hope, S.J., De Oliveira, C.M., Regis, L., Silva Filha, M.H.,
Dancer, B.N., Berry, C., 2006. Conjugal transfer of a toxin-coding megaplasmid from
Bacillus thuringiensis subsp. israelensis to mosquitocidal strains of Bacillus sphaericus. Appl.
Environ. Microbiol. 72, 17661770.
Geissbuhler, Y., Chaki, P., Emidi, B., Govella, N.J., Shirima, R., Mayagaya, V., Mtasiwa, D.,
Mshinda, H., Fillinger, U., Lindsay, S.W., Kannady, K., De Castro, M.C., Tanner, M.,
Killeen, G.F., 2007. Interdependence of domestic malaria prevention measures and
mosquito-human interactions in urban Dar es Salaam, Tanzania. Malar. J. 6, 126.
Geissbuhler, Y., Kannady, K., Chaki, P.P., Emidi, B., Govella, N.J., Mayagaya, V.,
Kiama, M., Mtasiwa, D., Mshinda, H., Lindsay, S.W., Tanner, M., Fillinger, U., De
Castro, M.C., Killeen, G.F., 2009. Microbial larvicide application by a large-scale,
community-based program reduces malaria infection prevalence in urban Dar es Salaam,
Tanzania. PLoS One 4, e5107.
Georghiou, G.P., Lagunes-Tejeda, A., 1991. The Occurrence of Resistance to Pesticides in
Arthropods. Food and Agriculture Organization of the United Nations. Report FAO-
AGP-MISC/91-1, 318 p.
Giraldo-Calderon, G.I., Perez, M., Morales, C.A., Ocampo, C.B., 2008. Evaluation of the
triflumuron and the mixture of Bacillus thuringiensis plus Bacillus sphaericus for control of
the immature stages of Aedes aegypti and Culex quinquefasciatus (Diptera: Culicidae) in
catch basins. Biomedica 28, 224233.
Goldberg, L.H., Margalit, J., 1978. A bacterial spore demonstrating rapid larvicidal activity
against Anopheles segentii, Uranotaenia unguiculata, Culex univitatus, Aedes aegypti and Culex
pipiens. Mosq. News 37, 355358.
Gonzalez, C.I., Bhattacharya, A., Wang, W., Peltz, S.W., 2001. Nonsense-mediated mRNA
decay in Saccharomyces cerevisiae. Gene 274, 1525.
Gould, F., Anderson, A., Jones, A., Sumerford, D., Heckel, D.G., Lopez, J., Micinski, S.,
Leonard, R., Laster, M., 1997. Initial frequency of alleles for resistance to Bacillus thur-
ingiensis toxins in field populations of Heliothis virescens. Proc. Natl. Acad. Sci. U.S.A
94, 35193523.
Groves, R.L., Meisch, M.V., 1996. Laboratory and field plot bioassay of Bacillus sphaericus
against Arkansas mosquito species. J. Am. Mosq. Control Assoc. 12, 220224.
Guerchicoff, A., Ugalde, R.A., Rubinstein, C.P., 1997. Identification and characterization
of a previously undescribed cyt gene in Bacillus thuringiensis subsp. israelensis. Appl. Envi-
ron. Microbiol. 63, 27162721.
Guidi, V., Patocchi, N., Luthy, P., Tonolla, M., 2011. Distribution of Bacillus thuringiensis
subsp. israelensis in soil of a swiss wetland reserve after 22 years of mosquito control. Appl.
Environ. Microbiol. 77, 36633668.
Microbial Toxins for Mosquito Control 159
Guidi, V., Lehner, A., Luthy, P., Tonolla, M., 2013a. Dynamics of Bacillus thuringiensis var.
israelensis and Lysinibacillus sphaericus spores in urban catch basins after simultaneous appli-
cation against mosquito larvae. PLoS One 8, e55658.
Guidi, V., Luthy, P., Tonolla, M., 2013b. Comparison between diflubenzuron and a Bacillus
thuringiensis israelensis- and Lysinibacillus sphaericus-based formulation for the control of
mosquito larvae in urban catch basins in Switzerland. J. Am. Mosq. Control Assoc.
29, 138145.
Guillet, P., Kurtak, D.C., Phillipon, B., Meyer, R., 1990. Use of Bacillus thuringiensis for
onchorcercosis control in West Africa. In: de Barjac, H., Sutherland, D. (Eds.), Bacterial
Control of Mosquitoes and Black-Flies, first ed. Rutgers University Press, New
Brunswick, pp. 187201.
Guo, Q.Y., Cai, Q.X., Yan, J.P., Hu, X.M., Zheng, D.S., Yuan, Z.M., 2013. Single nucle-
otide deletion of cqm1 gene results in the development of resistance to Bacillus sphaericus
in Culex quinquefasciatus. J. Insect Physiol. 59, 967973.
Harbison, J.E., Henry, M., Xamplas, C., Berry, R., 2013. Experimental use of Natular XRT
tablets in a North Shore suburb of Chicago, IL. J. Am. Mosq. Control Assoc. 29, 237242.
Hazes, B., Read, R.J., 1995. A mosquitocidal toxin with a ricin-like cell-binding domain.
Nat. Struct. Biol. 2, 358359.
Hemingway, J., Hawkes, N.J., Mccarroll, L., Ranson, H., 2004. The molecular basis of
insecticide resistance in mosquitoes. Insect Biochem. Mol. Biol. 34, 653665.
Hertlein, B.C., Levy, R., Miller, T.W.J., 1979. Recycling and selective retrieval of Bacillus
sphaericus from soil in a mosquito habitat. J. Invertebr. Pathol. 15, 1520.
Hertlein, M.B., Mavrotas, C., Jousseaume, C., Lysandrou, M., Thompson, G.D., Jany, W.,
Ritchie, S.A., 2010. A review of spinosad as a natural product for larval mosquito control.
J. Am. Mosq. Control Assoc. 26, 6787.
Hindley, J., Berry, C., 1987. Identification, cloning and sequence analysis of the Bacillus
sphaericus 1593 41.9 kD larvicidal toxin gene. Mol. Microbiol. 1, 187194.
Hire, R.S., Hadapad, A.B., Dongre, T.K., Kumar, V., 2009. Purification and characterization
of mosquitocidal Bacillus sphaericus BinA protein. J. Invertebr. Pathol. 101, 106111.
Hire, R.S., Hadapad, A.B., Vijayalakshmi, N., Dongre, T.K., 2010. Characterization of
highly toxic indigenous strains of mosquitocidal organism Bacillus sphaericus. FEMS
Microbiol. Lett. 305, 155161.
Hougard, J.M., 1990. Formulations and persistence of Bacillus sphaericus in Culex quin-
quefasciatus larval sites in tropical Africa. In: De Barjac, H., Sutherland, D. (Eds.), Bacterial
Control of Mosquitoes and Black-Flies: Biochemistry, Genetics and Applications of
Bacillus thuringiensis Israelensis and Bacillus sphaericus. Rutgers University Press, New
Brunswick, NJ, pp. 295306.
Hougard, J.M., Back, C., 1992. Perspectives on the bacterial control of vectors in the tropics.
Parasitol. Today 8, 364366.
Hougard, J.M., Mbentengam, R., Lochouarn, L., Escaffre, H., Darriet, F., Barbazan, P.,
Quillevere, D., 1993. Campaign against Culex quinquefasciatus using Bacillus sphaericus:
results of a pilot project in a large urban area of equatorial Africa. Bull. WHO
71, 367375.
Hu, X., Li, J., Hansen, B.M., Yuan, Z., 2008. Phylogenetic analysis and heterologous expres-
sion of surface layer protein SlpC of Bacillus sphaericus C3-41. Biosci. Biotechnol. Bio-
chem. 72, 12571263.
Hua, G., Zhang, R., Abdullah, M.A., Adang, M.J., 2008. Anopheles gambiae cadherin
AgCad1 binds the Cry4Ba toxin of Bacillus thuringiensis israelensis and a fragment of
AgCad1 synergizes toxicity. Biochemistry 47, 51015110.
Hua, G., Zhang, R., Bayyareddy, K., Adang, M.J., 2009. Anopheles gambiae alkaline phospha-
tase is a functional receptor of Bacillus thuringiensis jegathesan Cry11Ba toxin.
Biochemistry 48, 97859793.
160 Maria Helena Neves Lobo Silva Filha et al.
Humphreys, M.J., Berry, C., 1998. Variants of the Bacillus sphaericus binary toxins: implica-
tions for differential toxicity of strains. J. Invertebr. Pathol. 71, 184185.
Hurst, T.P., Brown, M.D., Kay, B.H., Ryan, P.A., 2006. Evaluation of Melanotaenia
duboulayi (Atheriniformes: Melanotaeniidae), Hypseleotris galli (Perciformes: Eleotridae),
and larvicide VectoLex WG (Bacillus sphaericus) for integrated control of Culex
annulirostris. J. Am. Mosq. Control Assoc. 22, 418425.
Janecek, S., 1997. Alpha-amylase family: molecular biology and evolution. Prog. Biophys.
Mol. Biol. 67, 6797.
Jiang, Y., Mulla, M.S., 2009. Laboratory and field evaluation of spinosad, a biorational natural
product, against larvae of Culex mosquitoes. J. Am. Mosq. Control Assoc. 25, 456466.
Jones, J.W., Weathersbee 3rd., A.A., Efird, P., Meisch, M.V., 1990. Evaluation of Bacillus
sphaericus 2362 against Culex quinquefasciatus in septic ditches. J. Am. Mosq. Control
Assoc. 6, 496499.
Jones, G.W., Nielsen-Leroux, C., Yang, Y., Yuan, Z., Dumas, V.F., Monnerat, R.G.,
Berry, C., 2007. A new Cry toxin with a unique two-component dependency from
Bacillus sphaericus. FASEB J. 21, 41124120.
Jones, G.W., Wirth, M.C., Monnerat, R.G., Berry, C., 2008. The Cry48Aa-Cry49Aa
binary toxin from Bacillus sphaericus exhibits highly restricted target specificity. Environ.
Microbiol. 10, 24182424.
Jurat-Fuentes, J.L., Adang, M.J., 2004. Characterization of a Cry1Ac-receptor alkaline phos-
phatase in susceptible and resistant Heliothis virescens larvae. Eur. J. Biochem. 271,
31273135.
Kahindi, S.C., Midega, J.T., Mwangangi, J.M., Kibe, L.W., Nzovu, J., Luethy, P., Githure, J.,
Mbogo, C.M., 2008. Efficacy of vectobac DT and culinexcombi against mosquito larvae in
unused swimming pools in Malindi, Kenya. J. Am. Mosq. Control Assoc. 24, 538542.
Kale, A., Hire, R.S., Hadapad, A.B., Dsouza, S.F., Kumar, V., 2013. Interaction between
mosquito-larvicidal Lysinibacillus sphaericus binary toxin components: analysis of complex
formation. Insect Biochem. Mol. Biol. 43, 10451054.
Kalfon, A., Charles, J.F., Bourgouin, C., de Barjac, H., 1984. Sporulation of Bacillus sphaericus
2297: an electron microscope study of crystal-like inclusion biogenesis and toxicity to
mosquito larvae. J. Gen. Microbiol. 130, 893900.
Karch, S., Charles, J.F., 1987. Toxicity, viability and ultrastructure of Bacillus sphaericus 2362
spore/crystal complex used in the field. Ann. Inst. Pasteur Microbiol. 138, 485492.
Karch, S., Coz, J., 1986. Recycling of Bacillus sphaericus in dead larvae of Culex pipiens
(Diptera: Culicidae). Cah. ORSTOM Entomol. Med. Parasitol. 24, 4143.
Karch, S., Monteny, N., Jullien, J.L., Sinegre, G., Coz, J., 1990. Control of Culex pipiens by
Bacillus sphaericus and role of nontarget arthropods in its recycling. J. Am. Mosq. Control
Assoc. 6, 4754.
Karch, S., Manzambi, Z.A., Salaun, J.J., 1991. Field trials with Vectolex (Bacillus sphaericus)
and Vectobac (Bacillus thuringiensis (H-14)) against Anopheles gambiae and Culex quin-
quefasciatus breeding in Zaire. J. Am. Mosq. Control Assoc. 7, 176179.
Karch, S., Asidi, N., Manzambi, Z.M., Salaun, J.J., 1992. Efficacy of Bacillus sphaericus against
the malaria vector Anopheles gambiae and other mosquitoes in swamps and rice fields in
Zaire. J. Am. Mosq. Control Assoc. 8, 376380.
Keiser, J., Maltese, M.F., Erlanger, T.E., Bos, R., Tanner, M., Singer, B.H., Utzinger, J.,
2005. Effect of irrigated rice agriculture on Japanese encephalitis, including challenges
and opportunities for integrated vector management. Acta Trop. 95, 4057.
Kellen, W.R., Clark, T.B., Lindegren, J.E., Ho, B.C., Rogoff, M.H., Singer, S., 1965. Bacil-
lus sphaericus Neide as a pathogen of mosquitoes. J. Invertebr. Pathol. 7, 442448.
Key, P.B., Scott, G.I., 1992. Acute toxicity of the mosquito larvicide, Bacillus sphaericus, to
the grass shrimp, Palaemonetes pugio, and mummichog, Fundulus heteroclitus. Bull. Envi-
ron. Contam. Toxicol. 49, 425430.
Microbial Toxins for Mosquito Control 161
Killeen, G.F., Smith, T.A., Ferguson, H.M., Mshinda, H., Abdulla, S., Lengeler, C.,
Kachur, S.P., 2007. Preventing childhood malaria in Africa by protecting adults from
mosquitoes with insecticide-treated nets. PLoS Med. 4, e229.
Knight, P.J., Crickmore, N., Ellar, D.J., 1994. The receptor for Bacillus thuringiensis CrylA(c)
delta-endotoxin in the brush border membrane of the lepidopteran Manduca sexta is ami-
nopeptidase N. Mol. Microbiol. 11, 429436.
Kramer, V.L., 1990. Efficacy and persistence of Bacillus sphaericus, Bacillus thuringiensis var.
israelensis, and methoprene against Culiseta incidens (Diptera: Culicidae) in tires. J. Econ.
Entomol. 83, 12801285.
Kramer, L.D., Styer, L.M., Ebel, G.D., 2008. A global perspective on the epidemiology of
West Nile virus. Annu. Rev. Entomol. 53, 6181.
Krasikov, V.V., Karelov, D.V., Firsov, L.M., 2001. Alpha-glucosidases. Biochemistry
(Mosc.) 66, 267281.
Krych, V.K., Johnson, J.L., Yousten, A.A., 1980. Deoxyribonucleic acid homologies among
strans of Bacillus sphaericus. Int. J. Syst. Bacteriol. 30, 253256.
Kumar, A., Sharma, V.P., Sumodan, P.K., Thavaselvam, D., Kamat, R.H., 1994. Malaria
control utilizing Bacillus sphaericus against Anopheles stephensi in Panaji, Goa. J. Am. Mosq.
Control Assoc. 10, 534539.
Kumar, A., Sharma, V.P., Thavaselvam, D., Sumodan, P.K., Kamat, R.H., Audi, S.S.,
Surve, B.N., 1996. Control of Culex quinquefasciatus with Bacillus sphaericus in Vasco City,
Goa. J. Am. Mosq. Control Assoc. 12, 409413.
Kumar, A.N., Murugan, K., Madhiyazhagan, P., Prabhu, K., 2011. Spinosad and neem seed
kernel extract as bio-controlling agents for malarial vector, Anopheles stephensi and non-
biting midge, Chironomus circumdatus. Asian Pac. J. Trop. Med. 4, 614618.
Kunthic, T., Promdonkoy, B., Srikhirin, T., Boonserm, P., 2011. Essential role of
tryptophan residues in toxicity of binary toxin from Bacillus sphaericus. BMB Rep.
44, 674679.
Kuppusamy, M., Hoti, S.L., Balaraman, K., 1987. Residual activity of briquette & alginate
formulations of Bacillus sphaericus against mosquito larvae. Indian J. Med. Res.
86, 591596.
Lacey, L., 2007. Bacillus thuringiensis serovariety israelensis and Bacillus sphaericus for mosquito
control. J. Am. Mosq. Control Assoc. 23, 133163.
Lacey, L.A., Lacey, C.M., 1990. The medical importance of riceland mosquitoes and their
control using alternatives to chemical insecticides. J. Am. Mosq. Control Assoc. Suppl.
2, 193.
Lacey, L., Mulla, M.S., 1990. Safety of Bacillus thuringiensis (H-14) and Bacillus sphaericus to
non-target organisms in the aquatic environment. In: Laird, M., Lacey, L.A.,
Davidson, E.W. (Eds.), Safety of Microbial Insecticides. CRC Press, Boca Raton, FL,
pp. 168188.
Lacey, L.A., Siegel, J.P., 2000. Safety and ecotoxicology of entomopathogenic bacteria.
In: Charles, J.F., Nielsen-Leroux, C., Delecluse, A. (Eds.), Entomopathogenic Bacteria:
From Laboratory to Field Application. Kluwer Academic Publishers, Dordrecht, The
Netherlands, pp. 253273.
Lacey, L.A., Urbina, M., Heitzman, C., 1984. Sustained release formulations of Bacillus
sphaericus and Bacillus thuringiensis (H-14) for the control of container breeding Culex
quinquefasciatus. Mosq. News 44, 2632.
Lacey, L.A., Heitzman, C.M., Meisch, M., Billodeaux, J., 1986. Beecomist-applied Bacillus
sphaericus for the control of riceland mosquitoes. J. Am. Mosq. Control Assoc.
2, 548551.
Lacey, L.A., Lacey, C.M., Peacock, B., Thiery, I., 1988a. Mosquito host range and field
activity of Bacillus sphaericus isolate 2297 (serotype 25). J. Am. Mosq. Control Assoc.
4, 5156.
162 Maria Helena Neves Lobo Silva Filha et al.
Lacey, L.A., Ross, D.H., Lacey, C.M., 1988b. Experimental formulations of Bacillus
sphaericus for the control of anopheline and culicine larvae. J. Indus. Microbiol. 3, 3947.
Lecadet, M.M., 1996. La lutte bacteriologique contre les insectes: une vielle histoire tres
actuelle. Ann. Inst. Pasteur/Actualit. 7, 207216.
Li, T., Sun, F., Yuan, Z., Zhang, Y., Yu, J., Pang, Y., 2000. Coexpression of cyt1Aa of Bacil-
lus thuringiensis subsp. israelensis with Bacillus sphaericus binary toxin gene in acrystalliferous
strain of B. thuringiensis. Curr. Microbiol. 40, 322326.
Li, J., Yang, L., Hu, X., Zheng, D., Yan, J., Yuan, Z., 2013. Nanoscale mono- and multi-
layer cylinder structures formed by recombinant S-layer proteins of mosquitocidal Bacil-
lus sphaericus C3-41. Appl. Microbiol. Biotechnol. 97, 72757283.
Likitvivatanavong, S., Chen, J., Evans, A.M., Bravo, A., Soberon, M., Gill, S.S., 2011. Mul-
tiple receptors as targets of Cry toxins in mosquitoes. J. Agric. Food Chem.
59, 28292838.
Limpanawat, S., Promdonkoy, B., Boonserm, P., 2009. The C-terminal domain of BinA is
responsible for Bacillus sphaericus binary toxin BinA-BinB interaction. Curr. Microbiol.
59, 509513.
Lingenfelser, A., Rydzanicz, K., Kaiser, A., Becker, N., 2010. Mosquito fauna and perspec-
tives for integrated control of urban vector-mosquito populations in Southern Benin
(West Africa). Ann. Agric. Environ. Med. 17, 4957.
Liu, J.W., Hindley, J., Porter, A.G., Priest, F.G., 1993. New high-toxicity mosquitocidal
strains of Bacillus sphaericus lacking a 100-kilodalton-toxin gene. Appl. Environ.
Microbiol. 59, 34703473.
Liu, J.W., Porter, A.G., Wee, B.Y., Thanabalu, T., 1996. New gene from nine Bacillus
sphaericus strains encoding highly conserved 35.8-kilodalton mosquitocidal toxins.
App. Environ. Microbiol. 62, 21742176.
Liu, H., Cupp, E.W., Micher, K.M., Guo, A., Liu, N., 2004. Insecticide resistance and cross-
resistance in Alabama and Florida strains of Culex quinquefasciatus. J. Med. Entomol.
41, 408413.
Lord, J.C., 1991. Sustained release pellets for control of Culex larvae with Bacillus sphaericus.
J. Am. Mosq. Control Assoc. 7, 560564.
Lozano, L.C., Ayala, J.A., Dussan, J., 2011. Lysinibacillus sphaericus S-layer protein toxicity
against Culex quinquefasciatus. Biotechnol. Lett. 33, 20372041.
Luo, K., Sangadala, S., Masson, L., Mazza, A., Brousseau, R., Adang, M.J., 1997. The
Heliothis virescens 170 kDa aminopeptidase functions as "receptor A" by mediating specific
Bacillus thuringiensis Cry1A delta-endotoxin binding and pore formation. Insect Bio-
chem. Mol. Biol. 27, 735743.
Luo, W., Liu, C., Zhang, R., He, J., Han, B., 2014. Anticancer activity of binary toxins from
Lysinibacillus sphaericus against human lung cancer cell line A549. Nat. Prod. Commun.
9, 107110.
Maciel, A., Rocha, A., Marzochi, K.B., Medeiros, Z., Carvalho, A.B., Regis, L., Souza, W.,
Lapa, T., Furtado, A., 1996. Epidemiological study of bancroftian filariasis in Recife,
northeastern Brazil. Mem. Inst. Oswaldo Cruz 91, 449455.
Maheu-Giroux, M., Castro, M.C., 2013. Impact of community-based larviciding on the
prevalence of malaria infection in Dar es Salaam, Tanzania. PLoS One 8, e71638.
Majambere, S., Lindsay, S.W., Green, C., Kandeh, B., Fillinger, U., 2007. Microbial larvi-
cides for malaria control in the Gambia. Malar. J. 6, 76.
Marcombe, S., Darriet, F., Agnew, P., Etienne, M., Yp-Tcha, M.M., Yebakima, A.,
Corbel, V., 2011. Field efficacy of new larvicide products for control of multi-resistant
Aedes aegypti populations in Martinique (French West Indies). Am. J. Trop. Med. Hyg.
84, 118126.
Margalit, J., Dean, D., 1985. The story of Bacillus thuringiensis var. israelensis (B.t.i.). J. Am.
Mosq. Control Assoc. 1, 17.
Microbial Toxins for Mosquito Control 163
Marina, C.F., Bond, J.G., Munoz, J., Valle, J., Chirino, N., Williams, T., 2012. Spinosad:
a biorational mosquito larvicide for use in car tires in southern Mexico. Parasit. Vectors
5, 95.
Marinotti, O., James, A.A., 1990. An -glucosidase in the salivary glands of the vector mos-
quito, Aedes aegypti. Insect Biochem. Mol. Biol. 20, 619623.
Marinotti, O., De Brito, M., Moreira, C.K., 1996. Apyrase and -glucosidase in the salivary
glands of Aedes albopictus. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 113, 675679.
Matanmi, B.A., Federici, B.A., Mulla, M.S., 1990. Fate and persistence of Bacillus sphaericus used as
a mosquito larvicide in dairy wastewater lagoons. J. Am. Mosq. Control Assoc. 6, 384389.
Maxwell, C.A., Mohammed, K., Kisumku, U., Curtis, C.F., 1999. Can vector control play a
useful supplementary role against bancroftian filariasis? Bull. WHO 77, 138143.
Melo-Santos, M.A., Araujo, A.P., Rios, E.M., Regis, L., 2009. Long lasting persistence
of Bacillus thuringiensis serovar. israelensis larvicidal activity in Aedes aegypti (Diptera:
Culicidae) breeding places is associated to bacteria recycling. Biol. Control 49, 186191.
Menon, K.K.R., Rao, A.S., Amonkar, S.V., 1982. Toxic activity and histopathological effects
of Bacillus sphaericus (ISPC-5) on mosquito larvae. Indian J. Exp. Biol. 20, 768772.
Merritt, R.W., Lessard, J.L., Wessell, K.J., Hernandez, O., Berg, M.B., Wallace, J.R.,
Novak, J.A., Ryan, J., Merritt, B.W., 2005. Lack of effects of Bacillus sphaericus
(Vectolex) on nontarget organisms in a mosquito-control program in southeastern
Wisconsin: a 3-year study. J. Am. Mosq. Control Assoc. 21, 201212.
Montero Lago, G., Diaz Perez, M., Marrero Figueroa, A., Castillo Gonzalez, F.A., 1991. The
pilot project results of applications of the biolarvicide Bacillus sphaericus 2362 on mosquito
breeding grounds of the town of Santa Cruz del Norte (La Habana Province). Rev.
Cubana Med. Trop. 43, 3944.
Moreno, J.E., Acevedo, P., Martnez, A., Sanchez, V., Petterson, L., 2010. Evaluacion de la
persistencia de una formulacion comercial de Bacillus sphaericus en criaderos naturales de
anofelinos vectores de malaria en estado Bolvar, Venezuela. Bol. Mal. Salud Amb.
50, 109117.
Moser, J.B., Ramirez, X., Gonzalez, J.E., Herrera, M., 2002. Evaluacion de la efectividad de
Bacillus sphaericus contra larvas de Anopheles aquasalis Curry (Diptera: Culicidae) en
criaderos naturales del estado Sucre, Venezuela. Entomotropica 17, 15.
Moser, J.B., Herrera, M., Rivas, J.G., Puentes, N., Caraballo, R., Valero, J., 2012. Field trials
on the efficacy and persistence of three formulations of Bacillus sphaericus against larvae of
Anopheles aquasalis Curry in mangroves of Marino municipality, Sucre state, Venezuela.
Bol. Mal. Salud Amb. 52, 6777.
Mulla, M.S., Darwazeh, H., Davidson, E.W., Dulmage, H.T., 1984a. Efficacy and persis-
tence of the microbial agent Bacillus sphaericus against mosquito larvae in organically
enriched habitats. Mosq. News 44, 166173.
Mulla, M.S., Darwazeh, H., Davidson, E.W., Dulmage, H.T., 1984b. Larvicidal activity and
field efficacy of Bacillus sphaericus strains against mosquito larvae and their safety for non
target organisms. Mosq. News 44, 336342.
Mulla, M.S., Darwazeh, H.A., Ede, L., Kennedy, B., Dulmage, H.T., 1985. Efficacy and
field evaluation of Bacillus thuringiensis (H-14) and B. sphaericus against floodwater mos-
quitoes in California. J. Am. Mosq. Control Assoc. 1, 310315.
Mulla, M.S., Axelrod, H., Darwazeh, H.A., Matanmi, B.A., 1988a. Efficacy and longevity of
Bacillus sphaericus 2362 formulations for control of mosquito larvae in dairy wastewater
lagoons. J. Am. Mosq. Control Assoc. 4, 448452.
Mulla, M.S., Darwazeh, H.A., Tietze, N.S., 1988b. Efficacy of Bacillus sphaericus 2362 for-
mulations against floodwater mosquitoes. J. Am. Mosq. Control Assoc. 4, 172174.
Mulla, M.S., Rodcharoen, J., Ngamsuk, W., Tawatsin, A., Pan-Urai, P., Thavara, U., 1997.
Field trials with Bacillus sphaericus formulations against polluted water mosquitoes in a
suburban area of Bangkok, Thailand. J. Am. Mosq. Control Assoc. 13, 297304.
164 Maria Helena Neves Lobo Silva Filha et al.
Mulla, M.S., Su, T., Thavara, U., Tawatsin, A., Ngamsuk, W., Pan-Urai, P., 1999. Efficacy
of new formulations of the microbial larvicide Bacillus sphaericus against polluted water
mosquitoes in Thailand. J. Vector Ecol. 24, 99110.
Mulla, M.S., Thavara, U., Tawatsin, A., Kong-Ngamsuk, W., Chompoosri, J., Su, T., 2001.
Mosquito larval control with Bacillus sphaericus: reduction in adult populations in
low-income communities in Nonthaburi Province, Thailand. J. Vector Ecol. 26, 221231.
Mulla, M.S., Thavara, U., Tawatsin, A., Chomposri, J., Su, T., 2003. Emergence of
resistance and resistance management in field populations of tropical Culex quin-
quefasciatus to the microbial control agent Bacillus sphaericus. J. Am. Mosq. Control Assoc.
19, 3946.
Mulligan, S.F., Schaefer, C.H., Miura, T., 1978. Laboratory and field evaluation of
Bacillus sphaericus as a mosquito (Diptera-Culicidae) control agent. J. Econ. Entomol.
71, 774777.
Mulligan, S.F., Schaefer, C.H., Wilder, W.H., 1980. Efficacy and persistence of Bacillus
sphaericus and Bacillus thuringiensis H.14 against mosquitos under laboratory and field con-
ditions. J. Econ. Entomol. 73, 684688.
Munhenga, G., Masendu, H.T., Brooke, B.D., Hunt, R.H., Koekemoer, L.K., 2008. Pyre-
throid resistance in the major malaria vector Anopheles arabiensis from Gwave, a malaria-
endemic area in Zimbabwe. Malar. J. 7, 247.
Mwangangi, J.M., Kahindi, S.C., Kibe, L.W., Nzovu, J.G., Luethy, P., Githure, J.I.,
Mbogo, C.M., 2011. Wide-scale application of Bti/Bs biolarvicide in different aquatic
habitat types in urban and peri-urban Malindi, Kenya. Parasitol. Res. 108, 13551363.
Myers, P., Yousten, A.A., Davidson, E.W., 1979. Comparative studies of the mosquito-
larval toxin of Bacillus sphaericus SSII-1 and 1593. Can. J. Microbiol. 25, 12271231.
Nakamura, L.K., 2000. Phylogeny of Bacillus sphaericus-like organisms. Int. J. Syst.
Evol. Microbiol. 50 (Pt. 5), 17151722.
Narasu, M.L., Gopinathan, K.P., 1986. Purification of larvicidal protein from Bacillus
sphaericus 1593. Biochem. Biophys. Res. Commun. 141, 756761.
Narasu, M.L., Gopinathan, K.P., 1988. Effect of Bacillus sphaericus 1593 toxin on choline ace-
tyl transferase and mitochondrial oxidative activities of the mosquito larvae. Indian J.
Biochem. Biophys. 25, 253256.
Neide, E., 1904. Botanische beschreibund einiger sporenbildenden bacterien. Zentralbl.
Bakteriol. Parasitenkd. Infektionskr. Hyg. Abt. 12 (Pt. 1), 111.
Nene, V., Wortman, J.R., Lawson, D., Haas, B., Kodira, C., Tu, Z.J., Loftus, B., Xi, Z.,
Megy, K., Grabherr, M., Ren, Q., Zdobnov, E.M., Lobo, N.F., Campbell, K.S.,
Brown, S.E., et al., 2007. Genome sequence of Aedes aegypti, a major arbovirus vector.
Science 316, 17181723.
Nicolas, L., Dossou-Yovo, J., 1987. Differential effects of Bacillus sphaericus strain 2362 on
Culex quinquefasciatus and its competitor Culex cinereus in West Africa. Med. Vet.
Entomol. 1, 2327.
Nicolas, L., Darriet, F., Hougard, J.M., 1987. Efficacy of Bacillus sphaericus 2362 against larvae
of Anopheles gambiae under laboratory and field conditions in West Africa. Med. Vet.
Entomol. 1, 157162.
Nicolas, L., Lecroisey, A., Charles, J.F., 1990. Role of the gut proteinases from mosquito
larvae in the mechanism of action and the specificity of the Bacillus sphaericus toxin.
Can. J. Microbiol. 36, 804807.
Nicolas, L., Nielsen-Leroux, C., Charles, J.F., Delecluse, A., 1993. Respective role of the
42- and 51-kDa components of the Bacillus sphaericus toxin overexpressed in Bacillus thur-
ingiensis. FEMS Microbiol. Lett. 106, 275280.
Nielsen-Leroux, C., Charles, J.F., 1992. Binding of Bacillus sphaericus binary toxin to a
specific receptor on midgut brush-border membranes from mosquito larvae. Eur. J.
Biochem. 210, 585590.
Microbial Toxins for Mosquito Control 165
Nielsen-Leroux, C., Charles, J.F., Thiery, I., Georghiou, G.P., 1995. Resistance in a labo-
ratory population of Culex quinquefasciatus (Diptera: Culicidae) to Bacillus sphaericus
binary toxin is due to a change in the receptor on midgut brush-border membranes.
Eur. J. Biochem. 228, 206210.
Nielsen-Leroux, C., Pasquier, F., Charles, J.F., Sinegre, G., Gaven, B., Pasteur, N., 1997.
Resistance to Bacillus sphaericus involves different mechanisms in Culex pipiens
(Diptera: Culicidae) larvae. J. Med. Entomol. 34, 321327.
Nielsen-Leroux, C., Rao, D.R., Murphy, J.R., Carron, A., Mani, T.R., Hamon, S.,
Mulla, M.S., 2001. Various levels of cross-resistance to Bacillus sphaericus strains in Culex
pipiens (Diptera: Culicidae) colonies resistant to B. sphaericus strain 2362. Appl. Environ.
Microbiol. 67, 50495054.
Nielsen-Leroux, C., Pasteur, N., Pretre, J., Charles, J.F., Sheikh, H.B., Chevillon, C.,
2002. High resistance to Bacillus sphaericus binary toxin in Culex pipiens (Diptera:
Culicidae): the complex situation of West Mediterranean countries. J. Med. Entomol.
39, 729735.
Nishiwaki, H., Nakashima, K., Ishida, C., Kawamura, T., Matsuda, K., 2007. Cloning,
functional characterization, and mode of action of a novel insecticidal pore-forming
toxin, sphaericolysin, produced by Bacillus sphaericus. Appl. Environ. Microbiol. 73,
34043411.
Oei, C., Hindley, J., Berry, C., 1990. An analysis of the genes encoding the 51.4- and 41.9-
kDa toxins of Bacillus sphaericus 2297 by deletion mutagenesis: the construction of fusion
proteins. FEMS Microbiol. Lett. 60, 265273.
Oei, C., Hindley, J., Berry, C., 1992. Binding of purified Bacillus sphaericus binary toxin and
its deletion derivatives to Culex quinquefasciatus gut: elucidation of functional binding
domains. J. Gen. Microbiol. 138, 15151526.
Oliveira, C.M., Silva-Filha, M.H., Nielsen-Leroux, C., Pei, G., Yuan, Z., Regis, L., 2004.
Inheritance and mechanism of resistance to Bacillus sphaericus in Culex quinquefasciatus
(Diptera: Culicidae) from China and Brazil. J. Med. Entomol. 41, 5864.
Opota, O., Charles, J.F., Warot, S., Pauron, D., Darboux, I., 2008. Identification and char-
acterization of the receptor for the Bacillus sphaericus binary toxin in the malaria vector
mosquito, Anopheles gambiae. Comp. Biochem. Physiol. B Biochem. Mol. Biol.
149, 419427.
Opota, O., Gauthier, N.C., Doye, A., Berry, C., Gounon, P., Lemichez, E., Pauron, D.,
2011. Bacillus sphaericus binary toxin elicits host cell autophagy as a response to intoxica-
tion. PLoS One 6, e14682.
Paing, M., Myat, M., Thu Sebastian, A.A., 1987. Laboratory evaluation of Bacillus sphaericus
15934 and preliminary field trials for control of Culex quinquefasciatus in septic tanks.
J. Commun. Dis. 19, 164167.
Park, H.W., Bideshi, D.K., Federici, B.A., 2003. Recombinant strain of Bacillus thuringiensis
producing Cyt1A, Cry11B, and the Bacillus sphaericus binary toxin. Appl. Environ.
Microbiol. 69, 13311334.
Park, H.W., Bideshi, D.K., Wirth, M.C., Johnson, J.J., Walton, W.E., Federici, B.A., 2005.
Recombinant larvicidal bacteria with markedly improved efficacy against Culex vectors
of West Nile virus. Am. J. Trop. Med. Hyg. 72, 732738.
Park, H.W., Bideshi, D.K., Federici, B.A., 2007. The 20-kDa protein of Bacillus thuringiensis
subsp. israelensis enhances Bacillus sphaericus 2362 bin toxin synthesis. Curr. Microbiol.
55, 119124.
Park, H.W., Tang, M., Sakano, Y., Federici, B.A., 2009. A 1.1-kilobase region downstream
of the bin operon in Bacillus sphaericus strain 2362 decreases bin yield and crystal size in
strain 2297. Appl. Environ. Microbiol. 75, 878881.
Partridge, M.R., Berry, C., 2002. Insecticidal activity of the Bacillus sphaericus Mtx1 toxin
against Chironomus riparus. J. Invertebr. Pathol. 79, 135136.
166 Maria Helena Neves Lobo Silva Filha et al.
Pauchet, Y., Luton, F., Castella, C., Charles, J.F., Romey, G., Pauron, D., 2005. Effects of a
mosquitocidal toxin on a mammalian epithelial cell line expressing its target receptor.
Cell. Microbiol. 7, 13351344.
Payne, J.M., Davidson, E.W., 1984. Insecticidal activity of the crystalline parasporal inclu-
sions and other components of the Bacillus sphaericus 1593 spore complex. J. Invertebr.
Pathol. 43, 383388.
Pei, G., Oliveira, C.M., Yuan, Z., Nielsen-Leroux, C., Silva-Filha, M.H., Yan, J., Regis, L.,
2002. A strain of Bacillus sphaericus causes slower development of resistance in Culex quin-
quefasciatus. Appl. Environ. Microbiol. 68, 30033009.
Pena, G., Miranda-Rios, J., De La Riva, G., Pardo-Lopez, L., Soberon, M., Bravo, A., 2006.
A Bacillus thuringiensis S-layer protein involved in toxicity against Epilachna varivestis
(Coleoptera: Coccinellidae). Appl. Environ. Microbiol. 72, 353360.
Penner, H., Wilamowski, A., 1996. Susceptibility of larvae of the sandfly Phlebotomus papatasi
(Diptera: Psychodidae) to Bacillus sphaericus. Bull. Entomol. Res. 86, 173175.
Perez, C., Fernandez, L.E., Sun, J., Folch, J.L., Gill, S.S., Soberon, 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, 1830318308.
Perez, C., Munoz-Garay, C., Portugal, L.C., Sanchez, J., Gill, S.S., Soberon, M., Bravo, A.,
2007. Bacillus thuringiensis ssp. israelensis Cyt1Aa enhances activity of Cry11Aa toxin by facil-
itating the formation of a pre-pore oligomeric structure. Cell. Microbiol. 9, 29312937.
Petersen, L.R., Fischer, M., 2012. Unpredictable and difficult to controlthe adolescence of
West Nile virus. N. Engl. J. Med. 367, 12811284.
Phannachet, K., Raksat, P., Limvuttegrijeerat, T., Promdonkoy, B., 2010. Production and
characterization of N- and C-terminally truncated Mtx2: a mosquitocidal toxin from
Bacillus sphaericus. Curr. Microbiol. 61, 549553.
Poncet, S., Delecluse, A., Anello, D., Klier, A., Rapoport, G., 1994. Transfer and expression
of the CryIVB and CryIVD genes of Bacillus thuringiensis subs. israelensis in Bacillus
sphaericus 2297. FEMS Microbiol. Lett. 117, 9196.
Poncet, S., Bernard, C., Dervyn, E., Cayley, J., Klier, A., Rapoport, G., 1997. Improvement
of Bacillus sphaericus toxicity against dipteran larvae by integration, via homologous
recombination, of the Cry11A toxin gene from Bacillus thuringiensis subsp. israelensis.
Appl. Environ. Microbiol. 63, 44134420.
Poophati, S., Mani, C., Vignesh, V., Praba, V.L., Thirugnanasambantham, K., 2013. Geno-
typic diversity of mosquitocidal bacteria (Bacillus sphaericus, B. thuringiensis, and B. cereus)
newly isolated from natural sources. Appl. Biochem. Biotechnol. 171, 22332246.
Porta, H., Cancino-Rodezno, A., Soberon, M., Bravo, A., 2011. Role of MAPK p38 in the
cellular responses to pore-forming toxins. Peptides 32, 601606.
Porter, A.G., Davidson, E.W., Liu, J.W., 1993. Mosquitocidal toxins of bacilli and their
genetic manipulation for effective biological control of mosquitoes. Microbiol. Rev.
57, 838861.
Prabhu, D.I., Sankar, S.G., Vasan, P.T., Piriya, P.S., Selvan, B.K., Vennison, S.J., 2013.
Molecular characterization of mosquitocidal Bacillus sphaericus isolated from Tamil Nadu,
India. Acta Trop. 127, 158164.
Pradeepkumar, N., Sabesan, S., Kuppusamy, M., Balaraman, K., 1988. Effect of controlled
release formulation of Bacillus sphaericus on Mansonia breeding. Indian J. Med. Res.
87, 1518.
Pridgeon, J.W., Pereira, R.M., Becnel, J.J., Allan, S.A., Clark, G.G., Linthicum, K.J., 2008.
Susceptibility of Aedes aegypti, Culex quinquefasciatus Say, and Anopheles quadrimaculatus
Say to 19 pesticides with different modes of action. J. Med. Entomol. 45, 8287.
Priest, F.G., Ebdrup, L., Zahner, V., Carter, P.E., 1997. Distribution and characterization of
mosquitocidal toxin genes in some strains of Bacillus sphaericus. Appl. Environ. Microbiol.
63, 11951198.
Microbial Toxins for Mosquito Control 167
Promdonkoy, B., Promdonkoy, P., Wongtawan, B., Wongtawan, B., Boonserm, P.,
Panyim, S., 2008. Cys31, Cys47, and Cys195 in BinA are essential for toxicity of a binary
toxin from Bacillus sphaericus. Curr. Microbiol. 56, 334338.
Ragoonanansingh, R.N., Njunwa, K.J., Curtis, C.F., Becker, N., 1992. A field study of
Bacillus sphaericus for the control of culicine and anopheline mosquito larvae in Tanzania.
Bull. Soc. Vector Ecol. 17, 4550.
Ramoska, W.A., Singer, S., Levy, R., 1977. Bioassay of three strains of Bacillus sphaericus on
field-collected mosquito larvae. J. Invertebr. Pathol. 30, 151154.
Ranson, H., Nguessan, R., Lines, J., Moiroux, N., Nkuni, Z., Corbel, V., 2011. Pyrethroid
resistance in African anopheline mosquitoes: what are the implications for malaria con-
trol? Trends Parasitol. 27, 9198.
Rao, D.R., Mani, T.R., Rajendran, R., Joseph, A.S., Gajanana, A., Reuben, R., 1995.
Development of a high level of resistance to Bacillus sphaericus in a field population of
Culex quinquefasciatus from Kochi, India. J. Am. Mosq. Control Assoc. 11, 15.
Regis, L.N., Nielsen-Leroux, C., 2000. Management of resistance to bacterial vector con-
trol. In: Charles, J.F., Nielsen-Leroux, C., Delecluse, A. (Eds.), Entomopathogenic Bac-
teria: From Laboratory to Field Application. Kluwer Academic Publisher, Dordrecht,
pp. 419439.
Regis, L., Silva-Filha, M.H., De Oliveira, C.M., Rios, E.M., Da Silva, S.B., Furtado, A.F.,
1995. Integrated control measures against Culex quinquefasciatus, the vector of filariasis in
Recife. Mem. Inst. Oswaldo Cruz 90, 115119.
Regis, L., Furtado, A.F., Oliveira, C.M., Bezerra, C.B., Silva, L.R., Araujo, J., Maciel, A.,
Silva-Filha, M.H., Silva, S.B., 1996. Integrated control of the filariasis vector with com-
munity participation in an urban area of Recife, Pernambuco, Brazil. Cad. Saude Publica
12, 473482.
Regis, L., Da Silva, S.B., Melo-Santos, M.A., 2000a. The use of bacterial larvicides in
mosquito and black fly control programmes in Brazil. Mem. Inst. Oswaldo Cruz
95 (Suppl. 1), 207210.
Regis, L., Oliveira, C.M., Silva-Filha, M.H., Silva, S.B., Maciel, A., Furtado, A.F., 2000b.
Efficacy of Bacillus sphaericus in control of the filariasis vector Culex quinquefasciatus in an
urban area of Olinda, Brazil. Trans. R. Soc. Trop. Med. Hyg. 94, 488492.
Regis, L., Silva-Filha, M.H., Nielsen-Leroux, C., Charles, J.F., 2001. Bacteriological larvi-
cides of dipteran disease vectors. Trends Parasitol. 17, 377380.
Reinert, D.J., Carpusca, I., Aktories, K., Schulz, G.E., 2006. Structure of the mosquitocidal
toxin from Bacillus sphaericus. J. Mol. Biol. 357, 12261236.
Rinkevich, F.D., Du, Y., Dong, K., 2013. Diversity and convergence of sodium
channel mutations involved in resistance to pyrethroids. Pestic. Biochem. Physiol.
106, 93100.
Rivera, P.L.E., Lopez, M., Valle, S., Delgado, M., Lopez, D., Larios, F., 1997. Evaluation de
la efetividad biolarvicida y residualidad de Bacillus sphaericus (cepa 2362) para el control de
Anopheles albimanus en la costa del Lago Xolotlan, Managua, Nicaragua. Rev. Nica.
Entomol. 42, 714.
Robert, L.L., Perich, M.J., Schlein, Y., Jacobson, R.L., Wirtz, R.A., Lawyer, P.G.,
Githure, J.I., 1997. Phlebotomine sand fly control using bait-fed adults to carry the lar-
vicide Bacillus sphaericus to the larval habitat. J. Am. Mosq. Control Assoc. 13, 140144.
Robert, L.L., Perich, M.J., Schlein, Y., Jacobson, J.L., 1998. Bacillus sphaericus inhibits hatch-
ing of phlebotomine sand fly eggs. J. Am. Mosq. Control Assoc. 14, 351352.
Robertson, J.L., Preisler, H.K., Ng, S.S., Hinkle, L.A., Gelernter, W.D., 1995. Natural var-
iations: a complicating factor in bioassays with chemical and microbial pesticides. J. Econ.
Entomol. 88, 110.
Rodcharoen, J., Mulla, M.S., 1994. Resistance development in Culex quinquefasciatus to the
microbial agent Bacillus sphaericus. J. Econ. Entomol. 87, 11331140.
168 Maria Helena Neves Lobo Silva Filha et al.
Rodcharoen, J., Mulla, M.S., 1996. Cross-resistance to Bacillus sphaericus strains in Culex quin-
quefasciatus. J. Am. Mosq. Control Assoc. 12, 247250.
Rodcharoen, J., Mulla, M.S., 1997. Biological fitness of Culex quinquefasciatus (Diptera:
Culicidae) susceptible and resistant to Bacillus sphaericus. J. Med. Entomol. 34, 510.
Rodcharoen, J., Mulla, M.S., Chaney, J.D., 1991. Microbial larvicides for the control of nui-
sance aquatic midges (Diptera: Chironomidae) inhabiting mesocosms and man-made
lakes in California. J. Am. Mosq. Control Assoc. 7, 5662.
Rodrigues, I.B., Tadei, W.P., Dias, J.M., 1998. Studies on the Bacillus sphaericus larvicidal activ-
ity against malarial vector species in Amazonia. Mem. Inst. Oswaldo Cruz 93, 441444.
Rodrigues, I.B., Tadei, W.P., Dias, J.M., 1999. Larvicidal activity of Bacillus sphaericus 2362
against Anopheles nuneztovari, Anopheles darlingi and Anopheles braziliensis (Diptera,
Culicidae). Rev. Inst. Med. Trop. Sao Paulo 41, 101105.
Rodrigues, I.B., Tadei, W.P., Santos, R.L.C., Santos, S., Baggio, J.B., 2008. Controle da
malaria: eficacia de formulados de Bacillus sphaericus 2362 contra especies de Anopheles
em criadouros artificiais-tanque de piscicultura e criadouros de olaria. Rev. Pat. Trop.
37, 161176.
Rojas, J.E., Mazzarri, M., Sojo, M., Garcia, A.G., 2001. Effectiveness of Bacillus sphaericus
strain 2362 on larvae of Anopheles nuneztovari. Invest. Clin. 42, 131146.
Romao, T.P., De Melo Chalegre, K.D., Key, S., Ayres, C.F., Fontes De Oliveira, C.M., De-
Melo-Neto, O.P., Silva-Filha, M.H., 2006. A second independent resistance mechanism
to Bacillus sphaericus binary toxin targets its alpha-glucosidase receptor in Culex quin-
quefasciatus. FEBS J. 273, 15561568.
Romao, T.P., De-Melo-Neto, O.P., Silva-Filha, M.H., 2011. The N-terminal third of the
BinB subunit from the Bacillus sphaericus binary toxin is sufficient for its interaction with
midgut receptors in Culex quinquefasciatus. FEMS Microbiol. Lett. 321, 167174.
Romi, R., Toma, L., Severini, F. Di, Luca, M., 2003. Susceptibility of Italian populations of
Aedes albopictus to temephos and to other insecticides. J. Am. Mosq. Control Assoc.
19, 419423.
Rungrod, A., Tjahaja, N.K., Soonsanga, S., Audtho, M., Promdonkoy, B., 2009. Bacillus
sphaericus Mtx1 and Mtx2 toxins co-expressed in Escherichia coli are synergistic against
Aedes aegypti larvae. Biotechnol. Lett. 31, 551555.
Rydzanicz, K., Lonc, E., Becker, N., 2009. Current procedures of the integrated urban
vector-mosquito control as an example in Cotonou (Benin, West Africa) and Wroclaw
area (Poland). Wiad. Parazytol. 55, 335340.
Saavedra-Rodriguez, K., Urdaneta-Marquez, L., Rajatileka, S., Moulton, M., Flores, A.E.,
Fernandez-Salas, I., Bisset, J., Rodriguez, M., Mccall, P.J., Donnelly, M.J., Ranson, H.,
Hemingway, J., Black, W.C.T., 2007. A mutation in the voltage-gated sodium channel
gene associated with pyrethroid resistance in Latin American Aedes aegypti. Insect Mol.
Biol. 16, 785798.
Sanitt, P., Promdonkoy, B., Boonserm, P., 2008. Targeted mutagenesis at charged residues in
Bacillus sphaericus BinA toxin affects mosquito-larvicidal activity. Curr. Microbiol.
57, 230234.
Sarkar, M., Borkotoki, A., Baruah, I., Bhattacharyya, I.K., Srivastava, R.B., 2009. Molecular
analysis of knock down resistance (kdr) mutation and distribution of kdr genotypes in a
wild population of Culex quinquefasciatus from India. Trop. Med. Int. Health
14, 10971104.
Schirmer, J., Just, I., Aktories, K., 2002a. The ADP-ribosylating mosquitocidal toxin from
Bacillus sphaericus: proteolytic activation, enzyme activity, and cytotoxic effects. J. Biol.
Chem. 277, 1194111948.
Schirmer, J., Wieden, H.J., Rodnina, M.V., Aktories, K., 2002b. Inactivation of the
elongation factor Tu by mosquitocidal toxin-catalyzed mono-ADP-ribosylation. Appl.
Environ. Microbiol. 68, 48944899.
Microbial Toxins for Mosquito Control 169
Schwartz, J.L., Potvin, L., Coux, F., Charles, J.F., Berry, C., Humphreys, M.J., Jones, A.F.,
Bernhart, I., Dalla Serra, M., Menestrina, G., 2001. Permeabilization of model lipid
membranes by Bacillus sphaericus mosquitocidal binary toxin and its individual compo-
nents. J. Membr. Biol. 184, 171183.
Schwede, T., Kopp, J., Guex, N., Peitsch, M.C., 2003. SWISS-MODEL: an automated pro-
tein homology-modeling server. Nucleic Acids Res. 31, 33813385.
Sebo, P., Bennardo, T., De La Torre, F., Szulmajster, J., 1990. Delineation of the minimal
portion of the Bacillus sphaericus 1593M toxin required for the expression of larvicidal
activity. Eur. J. Biochem. 194, 161165.
Servant, P., Rosso, M.L., Hamon, S., Poncet, S., Del Cluse, A., Rapoport, G., 1999. Pro-
duction of Cry11A and Cry11Ba toxins in Bacillus sphaericus confers toxicity towards
Aedes aegypti and resistant Culex populations. Appl. Environ. Microbiol. 65, 30213026.
Seyoum, A., Abate, D., 1997. Larvicidal efficacy of Bacillus thuringiensis var. israelensis and
Bacillus sphaericus on Anopheles arabiensis in Ethiopia. World J. Microbiol. Biotechnol.
13, 2124.
Shadduck, J.A., Singer, S., Lause, S., 1980. Lack of mammalian pathogenicity of entomocidal
isolates of Bacillus sphaericus. Environ. Entomol. 9, 403407.
Shanmugavelu, M., Rajamohan, F., Kathirvel, M., Elangovan, G., Dean, D.H.,
Jayaraman, K., 1998. Functional complementation of nontoxic mutant binary toxins
of Bacillus sphaericus 1593M generated by site-directed mutagenesis. Appl. Environ.
Microbiol. 64, 756759.
Shi, Y., Yuan, Z., Cai, Q., Yu, J., Yan, J., Pang, Y., 2001. Cloning and expression of the
binary toxin gene from Bacillus sphaericus IAB872 in a crystal-minus Bacillus thuringiensis
subsp. israelensis. Curr. Microbiol. 43, 2125.
Shililu, J.I., Tewolde, G.M., Brantly, E., Githure, J.I., Mbogo, C.M., Beier, J.C., Fusco, R.,
Novak, R.J., 2003. Efficacy of Bacillus thuringiensis israelensis, Bacillus sphaericus and temephos
for managing Anopheles larvae in Eritrea. J. Am. Mosq. Control Assoc. 19, 251258.
Siegel, J.P., Novak, R.J., 1997. Field trials of VectoLex CG, a Bacillus sphaericus larvicide, in
Illinois waste tires and catch basins. J. Am. Mosq. Control Assoc. 13, 305310.
Siegel, J.P., Novak, R.J., 1999. Duration of activity of the microbial larvicide VectoLex CG
(Bacillus sphaericus) in Illinois catch basins and waste tires. J. Am. Mosq. Control Assoc.
15, 366370.
Siegel, J.P., Shadduck, J.A., 1990. Mammalian safety of Bacillus sphaericus. In: De Barjac, H.,
Sutherland, D. (Eds.), Bacterial Control of Mosquitoes and Black Flies: Biochemistry,
Genetics, and Applications of Bacillus thuringiensis israelensis and Bacillus sphaericus.
Rutgers University Press, New Brunswick, NJ, pp. 202217.
Siegel, J.P., Smith, A.R., Novak, R.J., 2001. Recovery of commercially produced Bacillus
thuringiensis var. israelensis and Bacillus sphaericus from tires and prevalence of bacilli in arti-
ficial and natural containers. J. Am. Mosq. Control Assoc. 17, 3341.
Silapanuntakul, S., Pantuwatana, S., Bhumiratana, A., Charoensiri, K., 1983. The compar-
ative persistence of toxicity of Bacillus sphaericus strain 1593 and Bacillus thuringiensis sero-
type H-14 against mosquito larvae in different kinds of environments. J. Invertebr.
Pathol. 42, 387392.
Silva Filha, M.H.N.L., Peixoto, C.A., 2003. Immunocytochemical localization of the Bacillus
sphaericus toxin components in Culex quinquefasciatus (Diptera: Culicidae) larvae midgut.
Pestic. Biochem. Physiol. 77, 138146.
Silva-Filha, M.H., Regis, L., 1997. Reversal of low-level resistance to Bacillus sphaericus in a
field population of the southern house mosquito (Diptera:Culicidae) from an urban area
of Recife. Brazil. J. Econ. Entomol. 90, 299303.
Silva-Filha, M.H.N.L., Regis, L., Nielsen-Leroux, C., Charles, J.-F., 1995. Low level resis-
tance to Bacillus sphaericus in a field-treated population of Culex quinquefasciatus (Diptera:
Culicidae). J. Econ. Entomol. 88, 525530.
170 Maria Helena Neves Lobo Silva Filha et al.
Silva-Filha, M.H., Nielsen-Leroux, C., Charles, J.F., 1997. Binding kinetics of Bacillus
sphaericus binary toxin to midgut brush-border membranes of Anopheles and Culex sp.
mosquito larvae. Eur. J. Biochem. 247, 754761.
Silva-Filha, M.H., Nielsen-Leroux, C., Charles, J.F., 1999. Identification of the receptor for
Bacillus sphaericus crystal toxin in the brush border membrane of the mosquito Culex
pipiens (Diptera: Culicidae). Insect Biochem. Mol. Biol. 29, 711721.
Silva-Filha, M.H., Regis, L., Oliveira, C.M., Furtado, A.E., 2001. Impact of a 26-month
Bacillus sphaericus trial on the preimaginal density of Culex quinquefasciatus in an urban area
of Recife, Brazil. J. Am. Mosq. Control Assoc. 17, 4550.
Silva-Filha, M.H., Oliveira, C.M., Regis, L., Yuan, Z., Rico, C.M., Nielsen-Leroux, C., 2004.
Two Bacillus sphaericus binary toxins share the midgut receptor binding site: implications for
resistance of Culex pipiens complex (Diptera: Culicidae) larvae. FEMS Microbiol. Lett.
241, 185191.
Silva-Filha, M.H.N.L., Chalegre, K.D., Anastacio, D.B., Oliveira, C.M.F., Silva, S.B.,
Acioli, R.V., Hibi, S., Oliveira, D.C., Parodi, E.S.M., Marques Filho, C.A.M.,
Furtado, A.F., Regis, L., 2008. Culex quinquefasciatus field populations subjected to treat-
ment with Bacillus sphaericus did not display high resistance levels. Biol. Control
44, 227234.
Sinegre, G., Babinot, M., Quermel, J.M., Gavon, B., 1994. First field occurrence of Culex
pipiens resistance to Bacillus sphaericus in Southern France. In: Abstracts of the VIIIth
European Meeting of Vector Ecology. Barcelone, September 5-8 1994, p. 17.
Singer, S., 1973. Insecticidal activity of recent bacterial isolates and their toxins against mos-
quito larvae. Nature 244, 110111.
Singer, S., 1977. Isolation and development of bacterial pathogens in vectors. In: Biological
Regulation of Vectors. DHEW Publication No. (NIH) 771180, Bethesda, pp. 318.
Singer, S., 1980. Bacillus sphaericus for the control of mosquitoes. Biotechnol. Bioeng.
22, 13351355.
Singh, G.J., Gill, S.S., 1988. An electron microscope study of the toxic action of Bacillus
sphaericus in Culex quinquefasciatus larvae. J. Invertebr. Pathol. 52, 237247.
Singh, O.P., Bali, P., Hemingway, J., Subbarao, S.K., Dash, A.P., Adak, T., 2009. PCR-
based methods for the detection of L1014 kdr mutation in Anopheles culicifacies sensu lato.
Malar. J. 8, 154.
Singkhamanan, K., Promdonkoy, B., Chaisri, U., Boonserm, P., 2010. Identification of
amino acids required for receptor binding and toxicity of the Bacillus sphaericus binary
toxin. FEMS Microbiol. Lett. 303, 8491.
Singkhamanan, K., Promdonkoy, B., Srikhirin, T., Boonserm, P., 2013. Amino acid residues in
the N-terminal region of the BinB subunit of Lysinibacillus sphaericus binary toxin play a crit-
ical role during receptor binding and membrane insertion. J. Invertebr. Pathol. 114, 6570.
Skovmand, O., Bauduin, S., 1997. Efficacy of a granular formulation of Bacillus sphaericus
against Culex quinquefasciatus and Anopheles gambiae in West African countries. J. Vector
Ecol. 22, 4351.
Skovmand, O., Guillet, P., 2000. Sedimentation of Bacillus sphaericus in tap water and sewage
water. J. Invertebr. Pathol. 75, 243250.
Skovmand, O., Sanogo, E., 1999. Experimental formulations of Bacillus sphaericus and B. thur-
ingiensis israelensis against Culex quinquefasciatus and Anopheles gambiae (Diptera: Culicidae)
in Burkina Faso. J. Med. Entomol. 36, 6267.
Skovmand, O., Ouedraogo, T.D., Sanogo, E., Samuelsen, H., Toe, L.P., Baldet, T., 2009.
Impact of slow-release Bacillus sphaericus granules on mosquito populations followed in a
tropical urban environment. J. Med. Entomol. 46, 6776.
Skovmand, O., Ouedraogo, T.D., Sanogo, E., Samuelsen, H., Toe, L.P., Bosselmann, R.,
Czajkowski, T., Baldet, T., 2011. Cost of integrated vector control with improved san-
itation and road infrastructure coupled with the use of slow-release Bacillus sphaericus
granules in a tropical urban setting. J. Med. Entomol. 48, 813821.
Microbial Toxins for Mosquito Control 171
Smith, A.W., Camara-Artigas, A., Allen, J.P., 2004. Crystallization of the mosquito-
larvicidal binary toxin produced by Bacillus sphaericus. Acta Crystallogr. D Biol.
Crystallogr. 60, 952953.
Smith, A.W., Camara-Artigas, A., Brune, D.C., Allen, J.P., 2005. Implications of high-
molecular-weight oligomers of the binary toxin from Bacillus sphaericus. J. Invertebr. Pat-
hol. 88, 2733.
Souza-Neto, J.A., Machado, F.P., Lima, J.B., Valle, D., Ribolla, P.E., 2007. Sugar digestion
in mosquitoes: identification and characterization of three midgut alpha-glucosidases of
the neo-tropical malaria vector Anopheles aquasalis (Diptera: Culicidae). Comp. Bio-
chem. Physiol. A Mol. Integr. Physiol. 147, 9931000.
Sreshty, M.A., Kumar, K.P., Murty, U.S., 2011. Synergism between wild-type Bacillus thur-
ingiensis subsp. israelensis and B. sphaericus strains: a study based on isobolographic analysis
and histopathology. Acta Trop. 118, 1420.
Srisucharitpanit, K., Inchana, P., Rungrod, A., Promdonkoy, B., Boonserm, P., 2012.
Expression and purification of the active soluble form of Bacillus sphaericus binary toxin
for structural analysis. Protein Expr. Purif. 82, 368372.
Srisucharitpanit, K., Yao, M., Chimnaronk, S., Promdonkoy, B., Tanaka, I., Boonserm, P.,
2013. Crystallization and preliminary X-ray crystallographic analysis of the functional
form of BinB binary toxin from Bacillus sphaericus. Acta Crystallogr. F Struct. Biol. Crys-
tal. Commun. 69, 170173.
Srisucharitpanit, K., Yao, M., Promdonkoy, B., Chimnaronk, S., Tanaka, I., Boonserm, P.,
2014. Crystal structure of BinB: a receptor binding component of the binary toxin
from Lysinibacillus sphaericus. Proteins. http://dx.doi.org/10.1002/prot.24636. [Epub
ahead of print].
Sternberg, M., Grue, C., Conquest, L., Grassley, J., King, K., 2012. Efficacy, fate, and poten-
tial effects on salmonids of mosquito larvicides in catch basins in Seattle, Washington.
J. Am. Mosq. Control Assoc. 28, 206218.
Su, T.S., 2008. Evaluation of water-soluble pouches of Bacillus sphaericus applied as prehatch
treatment against Culex mosquitoes in simulated catch basins. J. Am. Mosq. Control
Assoc. 24, 5460.
Su, T., Mulla, M.S., 1999. Field evaluation of new water-dispersible granular formulations of
Bacillus thuringiensis ssp. israelensis and Bacillus sphaericus against Culex mosquitoes in
microcosms. J. Am. Mosq. Control Assoc. 15, 356365.
Su, T., Mulla, M.S., 2004. Documentation of high-level Bacillus sphaericus 2362 resistance in
field populations of Culex quinquefasciatus breeding in polluted water in Thailand. J. Am.
Mosq. Control Assoc. 20, 405411.
Suarez, M.F., Morales, C.A., 1999. Impact of Bacillus sphaericus (Vectolex(r) CG) in the con-
trol of Anopheles albimanus and Culex spp. in Buenaventura, Colombia. J. Am. Mosq.
Control Assoc. 15, 407415.
Sun, M., Luo, X., Dai, J., Qu, K., Liu, Z., Yu, L., Chen, Y., Yu, Z., 1996. Evaluation of
Bacillus thuringiensis and Bacillus sphaericus strains from chinese soils toxic to mosquito lar-
vae. J. Invertebr. Pathol. 68, 7477.
Sundararaj, R., Reuben, R., 1991. Evaluation of a microgel droplet formulation of Bacillus
sphaericus 1593 M (Biocide-S) for control of mosquito larvae in rice fields in southern
India. J. Am. Mosq. Control Assoc. 7, 556559.
Tabashnik, B.E., Finson, N., Johnson, M.W., Heckel, D.G., 1994. Cross-Resistance to Bacil-
lus thuringiensis Toxin CryIF in the Diamondback Moth (Plutella xylostella). Appl. Envi-
ron. Microbiol. 60, 46274629.
Tabashnik, B.E., Carriere, Y., Dennehy, T.J., Morin, S., Sisterson, M.S., Roush, R.T.,
Shelton, A.M., Zhao, J.Z., 2003. Insect resistance to transgenic Bt crops: lessons from
the laboratory and field. J. Econ. Entomol. 96, 10311038.
Tabashnik, B.E., Fabrick, J.A., Henderson, S., Biggs, R.W., Yafuso, C.M., Nyboer, M.E.,
Manhardt, N.M., Coughlin, L.A., Sollome, J., Carriere, Y., Dennehy, T.J., Morin, S.,
172 Maria Helena Neves Lobo Silva Filha et al.
2006. DNA screening reveals pink bollworm resistance to Bt cotton remains rare after a
decade of exposure. J. Econ. Entomol. 99, 15251530.
Tanapongpipat, S., Luxananil, P., Promdonkoy, B., Chewawiwat, N., Audtho, M.,
Panyim, S., 2003. A plasmid encoding a combination of mosquito-larvicidal genes from
Bacillus thuringiensis subsp. israelensis and Bacillus sphaericus confers toxicity against a broad
range of mosquito larvae when expressed in Gram-negative bacteria. FEMS Microbiol.
Lett. 228, 259263.
Tandeau De Marsac, N., De La Torre, F., Szulmajster, J., 1987. Expression of the larvicidal
gene of Bacillus sphaericus 1593M in the cyanobacterium Anacystis nidulans R2. Mol. Gen.
Genet. 209, 396398.
Tangsongcharoen, C., Boonserm, P., Promdonkoy, B., 2011. Functional characterization of
truncated fragments of Bacillus sphaericus binary toxin BinB. J. Invertebr. Pathol.
106, 230235.
Tchicaya, E.S., Koudou, B.G., Keiser, J., Adja, A.M., Cisse, G., Tanner, M., Tano, Y.,
Utzinger, J., 2009. Effect of repeated application of microbial larvicides on malaria trans-
mission in central Cote dIvoire. J. Am. Mosq. Control Assoc. 25, 382385.
Tetreau, G., Patil, C.D., Chandor-Proust, A., Salunke, B.K., Patil, S.V., Despres, L., 2013.
Production of the bioinsecticide Bacillus thuringiensis subsp. israelensis with deltamethrin
increases toxicity towards mosquito larvae. Lett. Appl. Microbiol. 57, 151156.
Thanabalu, T., 1992. Cloning and characterization of a gene encoding a 100 kDa toxin from
Bacillus sphaericus SSII-1 and expression of insecticidal toxins in Caulobacter crescentus.
PhD, National University of Singapore.
Thanabalu, T., Porter, A.G., 1995. Efficient expression of a 100-kilodalton mosquitocidal
toxin in protease-deficient recombinant Bacillus sphaericus. Appl. Environ. Microbiol.
61, 40314036.
Thanabalu, T., Porter, A.G., 1996. A Bacillus sphaericus gene encoding a novel type of
mosquitocidal toxin of 31.8 kDa. Gene 170, 8589.
Thanabalu, T., Hindley, J., Jackson-Yap, J., Berry, C., 1991. Cloning, sequencing, and
expression of a gene encoding a 100-kilodalton mosquitocidal toxin from Bacillus
sphaericus SSII-1. J. Bacteriol. 173, 27762785.
Thanabalu, T., Hindley, J., Berry, C., 1992a. Proteolytic processing of the mosquitocidal
toxin from Bacillus sphaericus SSII-1. J. Bacteriol. 174, 50515056.
Thanabalu, T., Hindley, J., Brenner, S., Oei, C., Berry, C., 1992b. Expression of the
mosquitocidal toxins of Bacillus sphaericus and Bacillus thuringiensis subsp. israelensis by
recombinant Caulobacter crescentus, a vehicle for biological control of aquatic insect larvae.
Appl. Environ. Microbiol. 58, 905910.
Thanabalu, T., Berry, C., Hindley, J., 1993. Cytotoxicity and ADP-ribosylating activity of
the mosquitocidal toxin from Bacillus sphaericus SSII-1: possible roles of the 27- and
70-kilodalton peptides. J. Bacteriol. 175, 23142320.
Thiery, I., de Barjac, H., 1989. Selection of the most potent Bacillus sphaericus strains based on
activity ratios determined on 3 mosquito species. Appl. Microbiol. Biotechnol.
31, 577581.
Thiery, I., Back, C., Barbazan, P., Sinegre, G., 1996. Applications de Bacillus thuringiensis et
de B. sphaericus dans la demoustication et la lutte contre les vecteurs de maladies
tropicales. Ann. Institut Pasteur/Actualit. 7, 247260.
Thiery, I., Hamon, S., Delecluse, A., Orduz, S., 1998. The introduction into Bacillus
sphaericus of the Bacillus thuringiensis subsp. medellin Cyt1Ab1 gene results in higher sus-
ceptibility of resistant mosquito larva populations to B. sphaericus. Appl. Environ.
Microbiol. 64, 39103916.
Thompson, J.D., Higgins, D.G., Gibson, T.J., 1994. CLUSTAL W: improving the
sensitivity of progressive multiple sequence alignment through sequence weighting,
position-specific gap penalties and weight matrix choice. Nucleic Acids Res.
22, 46734680.
Microbial Toxins for Mosquito Control 173
Thorne, L., Garduno, F., Thompson, T., Decker, D., Zounes, M., Wild, M., Walfield, A.M.,
Pollock, T.J., 1986. Structural similarity between the lepidoptera- and diptera-specific
insecticidal endotoxin genes of Bacillus thuringiensis subsp. "kurstaki" and "israelensis".
J. Bacteriol. 166, 801811.
Tietze, N.S., Hester, P.G., Shaffer, K.R., Prescott, S.J., Schreiber, E.T., 1994. Integrated
management of waste tire mosquitoes utilizing Mesocyclops longisetus (Copepoda:
Cyclopidae), Bacillus thuringiensis var. israelensis, Bacillus sphaericus, and methoprene.
J. Am. Mosq. Control Assoc. 10, 363373.
Townson, H., Nathan, M.B., Zaim, M., Guillet, P., Manga, L., Bos, R., Kindhauser, M.,
2005. Exploiting the potential of vector control for disease prevention. Bull. WHO
83, 942947.
Treiber, N., Reinert, D.J., Carpusca, I., Aktories, K., Schulz, G.E., 2008. Structure and
mode of action of a mosquitocidal holotoxin. J. Mol. Biol. 381, 150159.
Tripet, F., Wright, J., Lanzaro, G., 2006. A new high-performance PCR diagnostic for the
detection of pyrethroid knockdown resistance kdr in Anopheles gambiae. Am. J. Trop.
Med. Hyg. 74, 658662.
Trisrisook, M., Pantuwatana, S., Bhumiratana, A., Panbangred, W., 1990. Molecular cloning
of the 130-kilodalton mosquitocidal delta-endotoxin gene of Bacillus thuringiensis subsp.
israelensis in Bacillus sphaericus. Appl. Environ. Microbiol. 56, 17101716.
Vadlamudi, R.K., Weber, E., Ji, I., Ji, T.H., Bulla Jr., L.A., 1995. Cloning and expression of a
receptor for an insecticidal toxin of Bacillus thuringiensis. J. Biol. Chem. 270, 54905494.
Valaitis, A.P., Jenkins, J.L., Lee, M.K., Dean, D.H., Garner, K.J., 2001. Isolation and partial
characterization of gypsy moth BTR-270, an anionic brush border membrane glyc-
oconjugate that binds Bacillus thuringiensis Cry1A toxins with high affinity. Arch. Bio-
chem. Biophys. 46, 186200.
Vasquez, M.I., Violaris, M., Hadjivassilis, A., Wirth, M.C., 2009. Susceptibility of Culex
pipiens (Diptera: Culicidae) field populations in Cyprus to conventional organic insecti-
cides, Bacillus thuringiensis subsp. israelensis, and methoprene. J. Med. Entomol.
46, 881887.
Wagner, E., Lykke-Andersen, J., 2002. mRNA surveillance: the perfect persist. J. Cell Sci.
115, 30333038.
Wahba, M.M., 2000. The influence of Bacillus sphaericus on the biology and histology of
Phlebotomus papatasi. J. Egypt. Soc. Parasitol. 30, 315323.
Wati, M.R., Thanabalu, T., Porter, A.G., 1997. Gene from tropical Bacillus sphaericus
encoding a protease closely related to subtilisins from Antarctic bacilli. Biochim. Biophys.
Acta 1352, 5662.
Wei, S., Cai, Q., Cai, Y., Yuan, Z., 2007. Lack of cross-resistance to Mtx1 from Bacillus
sphaericus in B. sphaericus-resistant Culex quinquefasciatus (Diptera: Culicidae). Pest Manag.
Sci. 63, 190193.
Weiser, J., 1984. A mosquito-virulent Bacillus sphaericus in adult Simulium damnosum from
northern Nigeria. Zentralbl. Mikrobiol. 139, 5760.
Wermelinger, E.D., Zanuncio, J.C., Rangel, E.F., Cecon, P.R., Rabinovitch, L., 2000.
Toxicity of Bacillus species to larvae of Lutzomyia longipalpis (L. & N.) (Diptera:
Psychodidae: Phlebotominae). An. Soc. Entomol. Bras. 26, 609614.
White, J.P., Lotay, H.K., 1980. Minimal nutritional requirements of Bacillus sphaericus
NCTC9602 and 26 others strains of this species: the majority grow and sporulate with
acetate as sole major source of carbon. J. Gen. Microbiol. 118, 1319.
WHO, 1983. Integrated Vector Control: 7th Rep. WHO Exp. Comm.VCB, Technical
Reportys Series, No. 668.
WHO, 1985. Informal Consultation on the Development of Bacillus sphaericus as Microbial
Larvicide. WHO, Geneva, pp. 124, TDR/BCV/SPHAERICUS/85.3.
WHO, 1986. Resistance of Vectors and Reservoirs of Disease to Pesticides: World Health
Organization technical report series 737. pp. 187.
174 Maria Helena Neves Lobo Silva Filha et al.
WHO, 1993. Report on the Workshop on the Large-Scale Use of Bacillus sphaericus to Con-
trol Culex quinquefasciatus in Urban Environments. Centre Pasteur Cameroun, Maroua,
Cameroun.
WHO, 2004. Global Strategic Framework for Integrated Vector Management. WHO, p. 12,
Document WHO/CDS/CPE/PVC/2004.10.
Wickremesingue, R.S.B., Mendis, C.L., 1980. Bacillus sphaericus spores from Sri-Lanka
demostrating rapid larvicidal activity on Culex quinquefasciatus. Mosq. News 40, 387389.
Wilkinson, M.F., Shyui, A.B., 2002. RNA surveillance by nuclear scanning? Nat. Cell Biol.
4, E144E147.
Wirth, M.C., 2010. Mosquito resistance to bacterial larvicidal proteins. Open J. Toxicol.
3, 101115.
Wirth, M.C., Federici, B.A., Walton, W.E., 2000a. Cyt1A from Bacillus thuringiensis syn-
ergizes activity of Bacillus sphaericus against Aedes aegypti (Diptera: Culicidae). App. Envi-
ron. Microbiol. 66, 10931097.
Wirth, M.C., Georghiou, G.P., Malik, J.I., Abro, G.H., 2000b. Laboratory selection for
resistance to Bacillus sphaericus in Culex quinquefasciatus (Diptera: Culicidae) from
California, USA. J. Med. Entomol. 37, 534540.
Wirth, M.C., Walton, W.E., Federici, B.A., 2000c. Cyt1A from Bacillus thuringiensis restores
toxicity of Bacillus sphaericus against resistant Culex quinquefasciatus (Diptera: Culicidae).
J. Med. Entomol. 37, 401407.
Wirth, M.C., Delecluse, A., Walton, W.E., 2001a. Cyt1Ab1 and Cyt2Ba1 from Bacillus thur-
ingiensis subsp. medellin and B. thuringiensis subsp. israelensis Synergize Bacillus sphaericus
against Aedes aegypti and resistant Culex quinquefasciatus (Diptera: Culicidae). Appl. Envi-
ron. Microbiol. 67, 32803284.
Wirth, M.C., Ferrari, J.A., Georghiou, G.P., 2001b. Baseline susceptibility to bacterial insec-
ticides in populations of Culex pipiens complex (Diptera: Culicidae) from California and
from the Mediterranean Island of Cyprus. J. Econ. Entomol. 94, 920928.
Wirth, M.C., Jiannino, J.A., Federici, B.A., Walton, W.E., 2004. Synergy between toxins
of Bacillus thuringiensis subsp. israelensis and Bacillus sphaericus. J. Med. Entomol. 41, 935941.
Wirth, M.C., Jiannino, J.A., Federici, B.A., Walton, W.E., 2005. Evolution of resistance
toward Bacillus sphaericus or a mixture of B. sphaericus+Cyt1A from Bacillus thuringiensis,
in the mosquito, Culex quinquefasciatus (Diptera: Culicidae). J. Invertebr. Pathol.
88, 154162.
Wirth, M.C., Yang, Y., Walton, W.E., Federici, B.A., Berry, C., 2007. Mtx toxins synergize
Bacillus sphaericus and Cry11Aa against susceptible and insecticide-resistant Culex quin-
quefasciatus larvae. Appl. Environ. Microbiol. 73, 60666071.
Wirth, M.C., Berry, C., Walton, W.E., Federici, B.A., 2014. Mtx toxins from Lysinibacillus
sphaericus enhance mosquitocidal cry-toxin activity and suppress cry-resistance in Culex
quinquefasciatus. J. Invertebr. Pathol. 115, 6267.
Woodburn, M.A., Yousten, A.A., Hilu, K.H., 1995. Random amplified polymorphic DNA
fingerprinting of mosquito-pathogenic and nonpathogenic strains of Bacillus sphaericus.
Int. J. Syst. Bacteriol. 45, 212217.
Wraight, S.P., Molloy, D.P., Singer, S., 1987. Studies on the culicine mosquito host range of
Bacillus sphaericus and Bacillus thuringiensis var. israelensis with notes on the effects of tem-
perature and instar on bacterial efficacy. J. Invertebr. Pathol. 49, 291302.
Xu, Z., Liu, F., Chen, J., Huang, F., Andow, D.A., Wang, Y., Zhu, Y.C., Shen, J., 2009.
Using an F(2) screen to monitor frequency of resistance alleles to Bt cotton in field
populations of Helicoverpa armigera (Hubner) (Lepidoptera: Noctuidae). Pest Manag.
Sci. 65, 391397.
Yadav, R.S., Sharma, V.P., Upadhyay, A.K., 1997. Field trial of Bacillus sphaericus strain
B-101 (serotype H5a, 5b) against filariasis and Japanese encephalitis vectors in India.
J. Am. Mosq. Control Assoc. 13, 158163.
Microbial Toxins for Mosquito Control 175
Yang, Y., Chen, H., Wu, S., Xu, X., Wu, Y., 2006. Identification and molecular detection of
a deletion mutation responsible for a truncated cadherin of Helicoverpa armigera. Insect
Biochem. Mol. Biol. 36, 735740.
Yang, Y., Chen, H., Wu, Y., Wu, S., 2007a. Mutated cadherin alleles from a field population
of Helicoverpa armigera confer resistance to Bacillus thuringiensis toxin Cry1Ac. Appl. Envi-
ron. Microbiol. 73, 69396944.
Yang, Y., Wang, L., Gaviria, A., Yuan, Z., Berry, C., 2007b. Proteolytic stability of
insecticidal toxins expressed in recombinant bacilli. Appl. Environ. Microbiol. 73,
218225.
Yap, H.H., Tan, H.T., Yahaya, A.M., Baba, R., Chong, N.L., 1991. Small-scale field trials of
Bacillus sphaericus (strain 2362) formulations against Mansonia mosquitoes in Malaysia.
J. Am. Mosq. Control Assoc. 7, 2429.
Yap, W.H., Thanabalu, T., Porter, A.G., 1994. Expression of mosquitocidal toxin genes in a
gas-vacuolated strain of Ancylobacter aquaticus. Appl. Environ. Microbiol. 60, 41994202.
Yoshida, K., Hidaka, H., Miyado, S., Shibata, U., Saito, K., Yamada, Y., 1977. Purification
and some properties of Bacillus sphaericus protease. Agric. Biol. Chem. 41, 745754.
Young, M.D., Undeen, A.H., Dame, D.A., Wing, S.R., 1990. The effect of Bacillus
sphaericus upon the susceptibility of Anopheles quadrimaculatus to Plasmodium berghei.
J. Am. Mosq. Control Assoc. 6, 139140.
Yousten, A.A., 1984a. Bacillus sphaericus: microbiological factors related to its potential as a
mosquito larvicide. Adv. Biotechnol. Processes 3, 315343.
Yousten, A.A., 1984b. Bacteriophage typing of mosquito pathogenic strains of Bacillus
sphaericus. J. Invertebr. Pathol. 43, 124125.
Yousten, A.A., Davidson, E.W., 1982. Ultrastructural analysis of spores and parasporal crys-
tals formed by Bacillus sphaericus 2297. Appl. Environ. Microbiol. 44, 14491455.
Yousten, A.A., De Barjac, H., Hedrick, J., Cosmao Dumanoir, V., Myers, P., 1980. Com-
parison between bacteriophage typing and serotyping for the differentiation of Bacillus
sphaericus strains. Ann. Microbiol. (Paris) 131B, 297308.
Yousten, A.A., Benfield, E.F., Campbell, R.P., Foss, S.S., Genthner, F.J., 1991. Fate of Bacil-
lus sphaericus 2362 spores following ingestion by nontarget invertebrates. J. Invertebr.
Pathol. 58, 427435.
Yousten, A.A., Genthner, F.J., Benfield, E.F., 1992. Fate of Bacillus sphaericus and Bacillus
thuringiensis serovar israelensis in the aquatic environment. J. Am. Mosq. Control Assoc.
8, 143148.
Yousten, A.A., Benfield, E.F., Genthner, F.J., 1995. Bacillus sphaericus mosquito pathogens in
the aquatic environment. Mem. Inst. Oswaldo Cruz 90, 125129.
Yuan, Z., Neilsen-Leroux, C., Pasteur, N., Delecluse, A., Charles, J.F., Frutos, R., 1999.
Cloning and expression of the binary toxin genes of Bacillus sphaericus C3-41 in a crystal
minus B. thuringiensis subsp. israelensis. Wei Sheng Wu Xue Bao 39, 2935.
Yuan, Z.M., Zhang, Y.M., Liu, E.Y., 2000. High-level field resistance to Bacillus sphaericus
C3-41 in Culex quinquefasciatus from Southern China. Biocontrol Sci. Technol.
10, 4351.
Yuan, Z., Rang, C., Maroun, R.C., Juarez-Perez, V., Frutos, R., Pasteur, N., Vendrely, C.,
Charles, J.F., Nielsen-Leroux, C., 2001. Identification and molecular structural predic-
tion analysis of a toxicity determinant in the Bacillus sphaericus crystal larvicidal toxin. Eur.
J. Biochem. 268, 27512760.
Yuan, Z.M., Pei, G.F., Regis, L., Nielsen-Leroux, C., Cai, Q.X., 2003. Cross-resistance
between strains of Bacillus sphaericus but not B. thuringiensis israelensis in colonies of the
mosquito Culex quinquefasciatus. Med. Vet. Entomol. 17, 251256.
Zahiri, N.S., Mulla, M.S., 2003. Susceptibility profile of Culex quinquefasciatus (Diptera:
Culicidae) to Bacillus sphaericus on selection with rotation and mixture of B. sphaericus
and B. thuringiensis israelensis. J. Med. Entomol. 40, 672677.
176 Maria Helena Neves Lobo Silva Filha et al.
Zahiri, N.S., Su, T., Mulla, M.S., 2002. Strategies for the management of resistance in mos-
quitoes to the microbial control agent Bacillus sphaericus. J. Med. Entomol. 39, 513520.
Zhao, J., Jin, L., Yang, Y., Wu, Y., 2010. Diverse cadherin mutations conferring resistance to
Bacillus thuringiensis toxin Cry1Ac in Helicoverpa armigera. Insect Biochem. Mol. Biol.
40, 113118.
Zheng, L., Whang, L.H., Kumar, V., Kafatos, F.C., 1995. Two genes encoding midgut-
specific maltase-like polypeptides from Anopheles gambiae. Exp. Parasitol. 81, 272283.
CHAPTER FOUR
Contents
1. Introduction 178
2. Bt-Based Biopesticides 179
2.1 History of use of Bt for insect control 179
2.2 Biopesticides based on Bt 180
2.3 Molecular eraFirst cloned Bt insecticidal protein genes 181
2.4 Transconjugation, recombinant strains and alternative delivery systems
for Bt-based biopesticides 182
3. Discovery, Characterization and Development of Insecticidal Protein Genes
as Crop Traits 184
3.1 Diversity of Bt insecticidal proteins 184
3.2 Biological activity of Bt insecticidal proteins 185
3.3 Bt insecticidal protein structure and function: Cry proteins 187
3.4 Cry protein mechanism of action 188
3.5 Bt insecticidal protein structure and function: Cyt proteins 190
3.6 Bt insecticidal protein receptors 191
3.7 Mechanisms of resistance to Bt insecticidal proteins 191
4. Discovery and Development of Bt Crops 193
4.1 The discovery and development process 193
4.2 Gene discovery 194
4.3 First demonstrated success of Bt Cry GE plants 196
4.4 Transformation technologies 197
4.5 Introgression and testing 198
5. Regulation 198
5.1 Product identification and characterization 201
5.2 Human health assessment 201
5.3 Environmental effects 203
5.4 Considerations for stacks 206
5.5 Continued regulatory oversight of commercialized GE events 206
Abstract
Bacillus thuringiensis (Bt) is a ubiquitous, spore-forming soil bacterium that is well known
for production of insecticidal proteins that are active on a wide range of pest insects. The
potential of Bt to be used as an insecticide was recognized in the early twentieth century
and since that time many Bt-based biopesticides have been commercialized. The
advent of modern molecular biology tools made it possible to engineer plants to
express the genes coding for Bt insecticidal proteins as a safe, convenient and highly
effective means to protect plants from insect damage. The first Bt crop was commer-
cialized in 1995, and today Bt corn, cotton and soybean are cultivated on ca. 76 million
hectares in 27 countries. First generation products containing single Bt genes were
followed by broader spectrum products containing multiple Bt genes with the most
recent generation of products contain multiple Bt genes encoding proteins that target
the same pest(s) but with differences in their mechanism of action (i.e. gene pyramids)
as a means of increasing product durability. Developing Bt crops is a long and expensive
process that by recent estimates averages 13 years at a cost of $136 million. The process
of obtaining approvals by government regulatory agencies is among the most critical in
the later stages of the development process and represents ca. 25% of the total cost in
bringing a Bt crop to the market. Multiple factors drive the search for novel insect resis-
tance (IR) traits and Bt remains a significant focus of new IR trait discovery.
1. INTRODUCTION
Bacillus thuringiensis (Bt) is a ubiquitous, spore-forming soil bacterium
that is well known for production of parasporal crystalline inclusions during
the stationary phase of cell growth. These parasporal inclusions are comprised
of insecticidal proteins known as -endotoxins, including those classified as
Cry (crystalline) or Cyt (cytolytic) proteins. The parasporal crystalline
inclusions produced by Bt are composed of a diversity of proteins across dis-
tinct phylogenetic groups of sequences (Crickmore et al., 1998) and (http://
www.lifesci.sussex.ac.uk/home/Neil_Crickmore/Bt/). Collectively, Cry
Bt Crops 179
proteins are active on a wide range of insects including those among the
orders of Lepidoptera, Diptera and Coleoptera (van Frankenhuyzen,
2009). Bt also produces soluble insecticidal proteins during the cell vegetative
growth phase before the onset of sporulation that are named Vips (vegetative
insecticidal proteins) (Estruch et al., 1996; Warren, 1997; http://www.
lifesci.sussex.ac.uk/home/Neil_Crickmore/Bt/).
Various subspecies of Bt have historically been developed for use as foliar
applied biopesticides (Sanahuja et al., 2011) and have a long history of safe
use (Siegel, 2001). With the advent of modern molecular biology tools, it has
become possible to engineer plants to express the genes coding for Bt insec-
ticidal proteins as a safe, convenient and highly effective means to protect
plants from insect damage. The development of insect resistant crops has
rapidly progressed since the commercial introduction of Bt potato in
1995 and Bt corn and cotton in 1996 (http://www.epa.gov/oppbppd1/
biopesticides/pips/pip_list.htm). Today insect resistance (IR) traits based
on Bt proteins have achieved a high rate of world-wide adoption ( James,
2013). A current challenge for Bt trait seed producers is protecting the
long-term durability of Bt trait technology. Innovative insect resistance
management (IRM) strategies include the use of genetically engineered
(GE) crops containing combinations of Bt genes encoding novel insecticidal
proteins (i.e. pyramids). This chapter provides an overview of the history of
Bt biopesticides leading to Bt crop development, the success of Bt-based IR
traits and future prospects for Bt as a source of IR trait technology.
2. Bt-BASED BIOPESTICIDES
2.1. History of use of Bt for insect control
Bt has a long history of safe use as a biopesticide for insect control (Siegel,
2001). For an elegant review of the early historical events in the discovery
and development of insecticidal bacteria with significant attention directed
at Bt see Federici (2005). The bacterium that became known as Bt was first
reported in Japanese literature by Ishiwata (1901) during study of bacterial
disease of silkworms. Later, Berliner (1915) described a similar Bacillus bac-
terium that killed flour moths and named the organism B. thuringiensis for the
Thuringia region in Germany where the bacterial disease was discovered.
Research into the utility of Bt as an insecticide followed (Mattes, 1927)
and activity in field trials against the European corn borer, Ostrinia nubilalis
(Hubner), was reported in 1930 (Husz, 1930). This work led to the devel-
opment of a Bt product known as Sporeine that was commercialized in
180 Kenneth E. Narva et al.
the late 1930s (Federici, 2005). The potential for Bt to be used as an insec-
ticide became more widely appreciated years after these early studies. Pub-
lications by Hannay (1953) on the Bt parasporal crystal bodies and
demonstration that the parasporal crystals were capable of killing silkworms
(Angus, 1954)set the stage for an increase in research focused on developing
Bt as an insect control agent.
the antibiotic selectable marker resistant genes that might cause environ-
mental safety concerns (Baum et al., 1998). Several examples of recombinant
Bt strains are listed in Table 4.2.
Mycogen Corporation used a different approach to produce novel bio-
pesticides based on over-expression of cry genes in recombinant Pseudomonas
fluorescens (Gaertner et al., 1993). This gene expression system used recom-
binant DNA technology to express Cry proteins at high levels under high-
density cell culture fermentation conditions. The recombinant bacteria were
fixed in a proprietary treatment that rendered cells non-viable without
impacting the activity of the insecticidal proteins. The fixed,
120
Genes
100
Holotypes
80
Bt genes (n)
60
40
20
0
19
19
19 6
19
19 8
19 9
19 0
19
19 2
19
19 4
19
19 6
19 7
19
20
20 0
20 1
20 2
20
20 4
20
20 6
20
20
20 9
20 0
20 1
20
85
8
87
8
8
9
91
9
93
9
95
9
9
98
99
0
0
0
03
0
05
0
07
08
0
1
1
12
13
Figure 4.1 Discovery of Bt genes recorded on the Bt Toxin Nomenclature Website
maintained by the Bt toxin nomenclature committee (Crickmore et al., 2014). The total
number of new Cry, Cyt and Vip genes recognized by the committee in a given year is
shown as Genes. The total number of new gene classes (as defined by the committee)
recognized in a given year is shown as Holotypes.
Figure 4.2 Protein crystal structures of representative Bt insecticidal proteins. (A) Three
dimensional structure of Cry1Aa1 (PDB code: 1CIY), a three domain Cry protein.
(B) Three-dimensional structure of the cytolytic crystal protein Cyt2Aa (PDB code: 1CBY).
Table 4.3 Bt insecticidal protein crystal structures available in the Protein Data Bank
(PDB) (Website: http://www.rcsb.org/pdb/home/home.do)
PDB
Protein accession Structure Citation Year
Cyt2A1 1CBY Non-three domain Li et al. (1996) 1996
Cry1Aa1 1CIY Three domain Grochulski et al. (1995) 1995
Cry3Aa1 1DLC Three domain Li et al. (1991) 1991
Cry2Aa 1I5P Three domain Morse et al. (2001)
Cry3Bb1 1JI6 Three domain Galitsky et al. (2001) 2001
Cry4Ba 1W99 Three domain Boonserm et al. (2005) 2005
Cry4Aa 2C9K Three domain Boonserm et al. (2006) 2006
Cyt2Ba 2RCI Non-Three domain Cohen et al. (2008) 2008
Cry8Ea1 3EB7 Three domain Guo et al. (2009) 2009
Cyt1Aa 3RON Non-Three domain Cohen et al. (2011) 2011
Cry5Ba1 4D8M Three domain Hui et al. (2012) 2012
Cry34Ab1 4JOX Non-Three domain, unpublished 2014
binary with Cry35Ab1
Cry35Ab1 4JP0 Non-Three domain, unpublished 2014
binary with Cry34Ab1
events that lead to pore formation, can be summarized as follows. Cry pro-
teins are often produced as protoxins that are first solubilized in the insect
midgut and then proteolytically processed to yield smaller, activated poly-
peptides. The activated Cry proteins then bind to specific receptors on
the surface of insect midgut epithelial cells. Receptor binding is followed
by assembly of activated Cry proteins into pores that result in colloid osmotic
lysis of midgut cells due to an influx of solutes from the midgut lumen. Cell
lysis leads to disruption of the midgut epithelium and, ultimately, death of
the insect larva. This is often considered the classical model for Bt mech-
anism of action (Fig. 4.3). However, many details of this model remain unre-
solved. Two models have been researched in recent years that propose more
detailed mechanistic steps leading to insect death. These models are the
sequential binding model leading to pore formation (reviewed in
Soberon et al., 2009; Soberon et al., 2010) that builds on the classical pore
formation model and the signalling pathway model wherein Bt protein
190 Kenneth E. Narva et al.
Ingestion
Proteolysis
Insecticidal protein activated by
midgut proteases
Binding
Activated insecticidal proteins bind receptors
on the surface of midgut epithelial cells
Membrane insertion
Pore formation
Increased permeability
Loss of membrane function
Insect dealth
Figure 4.3 Schematic representation of the steps leading to pore formation and insect
death according the classical model of Bt mechanism of action (Vachon et al., 2012).
many more Bt genomes in the very near future. As of March 7, 2014, there
are 12 completed genomes publicly available on NCBI (http://www.ncbi.
nlm.nih.gov/genome/genomes/486). The challenges of data analysis and
identification of genes encoding new Bt insecticidal proteins are being
addressed (Ye et al., 2012). Recently, a pangenomic study of Bt was reported
(Fang et al., 2011) in which chromosomes and plasmids encoding Cry pro-
teins were sequenced to a high degree of coverage for seven Bt strains. The
pangenomic approach, which does not intend to assemble all genomes to
completion, but rather to interrogate sequence space across multiple Bt
strains, coupled with advances in the scale and throughput of insect bioas-
says, represents a powerful approach to identify genes encoding novel Bt
insecticidal proteins. One can surmise that the growing number of new pro-
tein sequences in the Bt nomenclature database are at least in part due to the
impact of next-generation sequencing (Fig. 4.1).
5. REGULATION
GE crops undergo comprehensive regulatory reviews for human
health and environmental safety by agencies throughout the world. Indeed,
GE crops and food receive far greater regulatory and scientific scrutiny than
Bt Crops 199
than purified transgene product (Tier 2). Again, if the test population is not
affected at realistic exposure levels, of if effects observed would be acceptable
(for example less than would occur with alternative pest control tools), addi-
tional testing is not warranted. If significant effects are seen in Tier 2, addi-
tional testing can be conducted with whole plants in a green house or field
cages (Tier 3). Such tests allow more realistic spatial processes to function
that may more accurately reflect actual exposure under field conditions.
Finally, if these lower tier studies indicate potential for unacceptable harm,
a field study may be warranted whereby natural populations are monitored
under the same conditions as the proposed environmental release (Tier 4).
Progressing through the tiers increases the ecological relevance of the study
to the actual proposed release but decreases the ability to detect effects due to
greater variability in the test system. Methods or guidance for testing many
non-target organisms at several of the tiers are available in published litera-
ture (e.g. Romeis et al., 2011) or from regulatory agencies (e.g. U.S.
Environmental Protection Agency, 2007).
Complementing such hazard testing, exposure analysis is accomplished
by measuring the expression of the GE protein in representative tissues of
the crop that are fed upon by herbivores. This can include leaf, stalk, pollen,
flowers and fruits, depending on the tissues that are consumed. Expression is
measured at several time points in the life cycle of the plant to provide a
comprehensive assessment of the potential exposure of non-target organ-
isms. Data are also generated on the environmental fate of the GE protein,
typically examining the rate of degradation of the protein in agricultural soils
(Shan, 2011).
It is reasonable to expect that some non-target species may be sensitive to
the GE protein, especially those that are phylogenetically related to the target
pest species. For example, larvae of the monarch butterfly and some other
Lepidoptera are known to be sensitive to Bt proteins in the Cry1 class, which
are targeted at lepidopteran pests. Similarly, larvae of certain Chrysomelidae
are known to be sensitive to Bt proteins in the Cry3 class. The risk to such
organisms is characterized by integrating their estimated sensitivity to high
end estimate of exposure levels. Usually, conservative assumptions are made
that over-estimate the sensitivity and over-estimate the exposure. If this
characterization indicates that there is not a very low likelihood of a harmful
effect to the population of the non-target organism, field studies may be
warranted to investigate whether the estimated effects actually occur under
field conditions.
Bt Crops 205
regulatory oversight. The EPA may require on-going studies on the envi-
ronmental effects of a Bt crop when grown on a commercial scale. Such
studies generally are confirmatory in nature, providing additional data on
exposure and effects of the Bt proteins. The European Food Safety Author-
ity requires technology providers to conduct on-going general surveillance
for changes in the agricultural ecosystem that may be attributable to the
release of GE crop.
However, such post-market monitoring is rarely scientifically justified.
The regulatory risk assessment prior to launch is in most cases sufficiently
thorough that unanticipated effects are known not to occur. General surveil-
lance is not hypothesis-driven, and collection of environmental data pro-
vides no information as to the cause of any changes, and whether such
changes are harmful or undesirable. Without a testable hypothesis, general
surveillance has little utility and is unlikely to identify environmental effects
resulting from the GE crop. Post market monitoring (PMM) is only
warranted when pre-market risk assessment identifies potentially unaccept-
able risks, and these risks can only be tested using large scale studies. In these
rare instances, post market monitoring can help determine actual levels of
harm and the efficacy of mitigation measures under the field conditions
reflective of commercialization. For additional information on PMM and
policy considerations, see FAO Expert Consultation on Genetically
Modified Organisms in Crop Production and Their Effects on the
Environment (2005).
Several regulatory agencies around the world require the technology
provider to implement resistance management programs that are designed
to slow the adaptation of target pest populations to GE Bt crops thereby
extending their utility and their benefits to the environment. Even where
these programs are not required, technology providers nevertheless will
implement measures to protect the durability of the products (Head and
Greenplate, 2012; MacIntosh, 2010).
crops in all geographies where they are grown. Resistance management pro-
grams are focused on the primary pest species that are of greatest importance
to the continued value of the Bt crop.
The primary tactic to delay resistance is the use of refuges, or host plants
that do not contain Bt genes and allow the persistence of susceptible pests.
Susceptible individuals are thereby able to mate with any resistant individuals
that emerge from the Bt crops and maintain susceptible alleles in the
population.
To be fully effective in delaying resistance development in a field pop-
ulation, refuges should produce sufficient insects to overwhelm any resistant
insectsa ratio of 500 susceptible to 1 resistant has been used as a rule of
thumb (U.S. Environmental Protection Agency, 2001b). The refuge should
be in sufficiently close proximity to Bt fields that normal insect dispersal will
promote mating between refuge-produced and Bt-produced insects. Adult
insect emergence from the refuge should occur at the same time as emer-
gence of resistant insects from Bt crops.
Different forms of refuge are used in resistance management programs.
Natural refuges can be composed of crop or non-crop host plants, often of
different species from the Bt crop, but nevertheless of sufficient abundance,
proximity and temporal overlap to promote mating of susceptible insects
with resistant insects from the Bt crop. Natural refuges consisting of crop
and non-crop hosts of H. virescens and H. zea provide the refuge for Bt cotton
in the south and southeastern United States (Gould et al., 2002; Gustafson
et al., 2006; Jackson et al., 2004) and for H. armigera in China (Qiao et al.,
2010). Structured refuges are specifically grown in association with Bt crops,
and consist of non-Bt varieties, usually of the same species as the Bt crop. The
recommended amount and layout of the refuge vary by pest species and
crop. For example, for O. nubilalis in the U.S. Corn Belt and single-gene
Bt maize hybrids, non-Bt maize must be on an area that is at least 20% of
the area of the Bt crop and the refuge must be planted within mile
(800 m) of each Bt field (U.S. Environmental Protection Agency,
2001a). For D. saccharalis in the Argentina corn belt and single-gene Bt
maize, the recommendation is for 10% refuge within 800 m of the Bt maize
field. For WCR in the U.S. Corn Belt and pyramided Bt maize, 5% refuge is
required which must be planted within or adjacent to the Bt maize field.
Recently, refuge provided as seed blends with Bt seeds that produce two
or more Bt proteins against each key target pest have been released to sim-
plify the refuge planting and management by growers and to ensure that the
required refuge is present (Onstad et al., 2011).
Bt Crops 209
7.1.1 Cry1Ab
Cry1Ab is one of the most well studied 3-domain Cry proteins. The cry1Ab1
gene and encoded protein were first cloned and characterized by Wabiko
et al. (1986) from Bt subspecies Berliner. Cry1Ab is produced as a protoxin
that is activated by proteases in the insect midgut and functions by the clas-
sical pore formation mechanism of action described in Section 3.4. Cry1Ab
is noted for broad spectrum lepidopteran activity that includes economically
important stalk boring pests such as Ostrinia spp. and Diatraea spp. (van
Frankenhuyzen, 2009). Much is known regarding functional aspects and
mechanisms of IR to Cry1Ab and cross-resistance with other Cry proteins
(Heckel et al., 2007; Schnepf et al., 1998).
7.1.2 Cry1Ac
Cry1Ac from Bt kurstaki strain HD73 was first described by Adang et al.
(1985). Cry1Ac is another example of a classical three domain insecticidal
Bt Crops 211
7.1.3 Cry1Fa2
Cry1Fa2 was discovered by Payne and Sick (1993). Cry1Fa2 displays
sequence characteristics of classical three domain Cry proteins. The lepidop-
teran pest spectrum of Cry1Fa2 is quite broad and includes high potency
against Spodoptera spp. (van Frankenhuyzen, 2009), an attribute that makes
Cry1Fa2 valuable in combination with Cry1A proteins with low potency on
Spodoptera spp.
7.1.4 Cry1A.105
Developed by Monsanto Co., Cry1A.105 is a chimeric protein comprising
parts of four domains from other Cry proteins previously used in GE
plants. The amino acid sequences of domains 1 and 2 are identical with
the respective domains from Cry1Ab and Cry1Ac proteins, domain 3 is
almost identical to the Cry1F protein, and the C-terminal domain is iden-
tical to the Cry1Ac protein (http://cera-gmc.org/index.php?actiongm_
crop_database&modeShowProd&dataMON89034). As a result, the
Cry1A.105 chimeric protein combines most of the insecticidal properties
displayed by the Cry1A and Cry1F proteins.
7.1.5 Cry2Ab
Cry2Ab is a three domain protein that diverges in primary sequence and
predicted structure from Cry1A proteins. Cry2Ab was the first dual spectrum
Cry protein described with activity against both lepidopteran and dipteran
insects (Widner and Whiteley, 1989). Cry2Ab is valued for broad spectrum
lepidopteran activity and the ability to control resistant insect populations
(summarized in Schnepf et al., 1998). Based on lack of shared midgut binding
sites (Hernandez-Rodriguez et al., 2008), Cry2Ab proteins are candidates for
pyramiding with Cry1 proteins in insect resistant crops. Cry2Ab is often
targeted to chloroplasts in order to increase expression levels and reduce
negative plant phenotypes in GE plants (Corbin and Romano, 2006).
212 Kenneth E. Narva et al.
7.1.6 Cry2Ae
Cry2Ae is related in sequence to Cry2Ab. Cry2Ae represents a different
mechanism of action from Cry1 proteins based on lack of shared binding
sites (Caccia et al., 2010; Gouffon et al., 2011). Cry2Ae, like Cry2Ab, is
a candidate for combining with Cry1 proteins in pyramids for resistance
management.
7.1.7 Vip3Aa
Vip3Aa is a soluble insecticidal protein expressed during the vegetative
growth phase of Bt growth, prior to sporulation (Estruch et al., 1996).
Vip3Aa is broadly active on lepidopteran pests of corn and cotton. Vip3Aa
is different in sequence and mechanism of action compared with Cry1 or
Cry2A proteins (Lee et al., 2006). As a result, Vip3Aa is useful in combina-
tion with Cry1 or Cry2 proteins to slow the development of resistant insect
populations.
Proteins developed in GE corn for control of WCR are mCry3Aa,
eCry3A.1ab, Cry3Bb and Cry34Ab1/35Ab1. These proteins and other
GE approaches for the control of WCR were recently reviewed by
Narva et al. (2013).
7.1.9 eCry3.1Ab
The eCry3.1Ab protein is a hybrid resulting from exchange of the domain 3
variable region from a lepidopteran-active toxin, Cry1Ab with domain 3
from Cry3Aa (Walters et al., 2010). The resulting protein, eCry3.1Ab,
has higher activity against WCR than Cry3Aa. Another interesting feature
of Cry3A.1Ab is that it binds WCR midgut brush border membrane vesicles
Bt Crops 213
7.1.10 Cry3Bb1
Cry3Bb1 is active on coleopteran pests such as L. decemlineata (Donovan
et al., 1992). Cry3Bb1 is a three domain Bt protein with structural similarity
to many other Cry proteins (Galitsky et al., 2001) and functions by forming
ion channels in membranes (Von Tersch et al., 1994). The version of
Cry3Bb1 expressed in events MON863 and MON88017 from Monsanto
is a modified protein with six amino acid residue changes compared with
the native sequence (Vaughn et al., 2005).
7.1.11 Cry34Ab1/Cry35Ab1
The second major class of Bt toxins developed for protection of WCR injury
in maize are the binary Bt crystal proteins Cry34Ab1 and Cry35Ab1
(Cry34Ab1/Cry35Ab1) that function together as oral intoxicants of
WCR larvae (Ellis et al., 2002; Herman et al., 2002). Cry34Ab1/Cry35Ab1
are structurally different from the Cry3-type proteins described above.
Cry34Ab1 is one example of a family of 14 kDa proteins that have no pro-
tein sequence homology beyond the Bt Cry34 group (Schnepf et al., 2005),
whereas Cry35Ab1 is a member of a family of 44 kDa Bt proteins that share
low sequence homology to Bt Cry36Aa1, Bacillus sphaericus mosquitocidal
binary proteins BinA and BinB and B. sphaericus mosquitocidal binary pro-
tein, Cry49Aa1 ( Jones et al., 2007). Crystal structures for Cry34Ab1 and
Cry35Ab1 were recently solved (Cry34Ab1 PDB accession 4JOX;
Cry35Ab1 PDB accession 4JP0; Kelker et al., 2014). Cry34Ab1/Cry35Ab1
appears to function by disrupting the WCR midgut epithelium. Consistent
with this observation, Cry34Ab1/Cry35Ab1 formed ion channels in
artificial lipid membranes (Masson et al., 2004). Li et al. (2013) recently
demonstrated Cry34Ab1/Cry35Ab1-specific binding to WCR midgut
BBMV and a lack of competitive binding between Cry34Ab1/Cry35Ab1
and the coleopteran-active proteins Cry3Aa, Cry6Aa and Cry8Ba. Lastly,
Cry34Ab1/Cry35Ab1 are not cross resistant with Cry3Bb1 insect resistant
traits; Gassmann et al. (2011) recently demonstrated that field-derived WCR
populations with reduced susceptibility to Cry3Bb1 corn are susceptible to
Cry34Ab1/Cry35Ab1 maize.
214 Kenneth E. Narva et al.
7.3.1 Bt corn
The first Bt corn events were approved for cultivation in the United States in
1995 and commercialized in 1996 targeting control of O. nubilalis and other
stalk boring Lepidoptera such as southwestern corn borer, Diatraea grand-
iosella Dyer and D. sacharralis (Archer et al., 2001; Buntin et al., 2004).
O. nubilalis had not been a target for insecticide applications so the rapid
adoption of Bt corn was based on the yield improvements created by con-
trolling pest populations that were below the economic threshold justifying
insecticide applications (Catangui, 2003; Shelton et al., 2002).
Events expressing Cry1Ab (176, MON810 and Bt11), Cry1Ac
(DBT418), Cry9C (CBH351), Cry1Fa (TC1507) and Vip3Aa (MIR162)
have all been commercialized (Table 4.5) and 4 (MON810, Bt11,
TC1507 and MIR162) remain in the market today (Table 4.6). Events
176 (Cry1Ab) and DBT418 (Cry1Ac) were discontinued based on the supe-
rior performance of MON810 and Bt11 (Buntin et al., 2004; Walker et al.,
2000). The removal of event 176 from commercial production was hastened
when it became the focus of concerns over the potential for impact on sus-
ceptible, non-target Lepidoptera consuming plant tissue onto which pollen
from Bt corn had fallen (Hellmich et al., 2001) (Sears et al., 2001). StarLink
(event CBH-351 expressing Cry9C) was approved for cultivation and for
use in animal feed and industrial non-food uses. It was removed from com-
mercial use when traces of the product were found in the food supply
(Fox, 2001).
MON810 and Bt11, the single Bt gene events expressing Cry1Ab in the
market today, provide a high level of control of O. nubilalis and other stalk
Table 4.5 Corn events approved for cultivation and commercialized
Year approved Non-IR
Developer Event name OECD unique identifier Bt protein(s) Pest spectrum (cultivationUSA) genesa
Syngenta 176 SYN-EV176-9 Cry1Ab Lepidoptera 1995 pat
Monsanto MON810 MON-00810-6 Cry1Ab Lepidoptera 1996 nptII
Syngenta Bt11 SYN-BT011-1 Cry1Ab Lepidoptera 1996 pat
Dekalb Genetics Corporation DBT418 DKB-89614-9 Cry1Ac Lepidoptera 1997 bar
b
Aventis CropScience CBH-351 ACS-ZM004-3 Cry9C Lepidoptera 1998 pat
Dow AgroSciences TC1507 DAS-01507-1 Cry1Fa Lepidoptera 2001 pat
DuPont Pioneer
Monsanto MON863 MON-00863-5 Cry3Bb1 Coleoptera 2003 nptII
Dow AgroSciences DAS-59122-7 DAS-59122-7 Cry34Ab1 Coleoptera 2005 pat
DuPont Pioneer Cry35Ab1
Monsanto MON88017 MON-88017-3 Cry3Bb1 Coleoptera 2005 cp4 epsps
Syngenta MIR604 SYN-IR604-5 mCry3A Coleoptera 2007 pmi
Monsanto MON89034 MON-89034-3 Cry1A.105 Lepidoptera 2008
Cry2Ab
Syngenta MIR162 SYN-IR162-4 Vip3Aa20 Lepidoptera 2010 pmi
Syngenta 5307 SYN-05307-1 eCry3.1Ab Coleoptera 2012 pmi
a
Non-IR genes pat: a selectable marker which confers tolerance to the herbicide glufosinate ammonium in plant tissue. nptII: a selectable marker which confers the ability
to metabolize the antibiotics neomycin and kanamycin in plant tissue. bar: a selectable marker which confers tolerance to the herbicide glufosinate ammonium in plant
tissue. cp4 epsps: a selectable marker confers tolerance to the herbicide glyphosate in plant tissue. pmi: a selectable marker which confers the ability to utilize mannose as a
carbon source in plant tissue.
b
Approved for environmental release and use as animal feed only.
Bt Crops 217
Figure 4.5 Insect-protected Bt corn. (A) Contrast of corn root systems after heavy pres-
sure from larval western corn rootworm. Top row is non-Bt plants showing severe root
damage. Bottom row is plants from the same genetic background as the plants in the
top row but containing event DAS-59122-7. (B) Contrast of ears after heavy pressure
from larval western bean cutworm. Top row is ears from non-Bt plants showing severe
damage. Bottom row is ears from the same genetic background as the plants in the top
row but containing event TC1507. Photo credit: Bader Rutter.
7.3.2 Bt cotton
Bt cotton was commercially introduced in the United States in 1996 and was
rapidly adopted due in large part to its effectiveness in controlling
H. virescens, a pest that had developed resistance to virtually all insecticides
available at the time (Blanco, 2012). Bt cotton was also introduced in
Australia in 1996 where the primary target was H. armigera (Downes and
Bt Crops 225
7.3.3 Bt soybean
MON87701 is the only Bt event that has been commercialized in soybean
(Table 4.9). Monsanto commercialized this event in Brazil under the trade
name Intacta RR2 PRO which is a breeding stack with MON89788, an
event conferring the Roundup Ready 2 Yield trait. Published studies
show that events expressing Cry1Ac in soybean are efficacious against
important North and South American soybean pests including
C. includens, A. gemmatalis, Crocidosema aporema (Walsingham), Rachiplusia
nu (Guenee) and the yellow woolybear, Spilosoma virginica (F.) (Macrae
et al., 2005; McPherson and MacRae, 2009). A more recent study by
Bernardi et al. (2014) using soybeans containing the events in Intacta
RR2 PRO showed that the relatively low intrinsic potency of Cry1Ac
on Spodoptera translated to poor control of S. frugiperda, Spodoptera cosmioides
Walker and Spodoptera eridania (Stoll); economically important pests of soy-
beans in Brazil.
Foliage feeding Lepidoptera frequently reach economically damaging
levels in South American soybean production, so Bt soybean potentially
delivers a clear benefit to growers (de Freitas Bueno et al., 2011). The value
226 Kenneth E. Narva et al.
Figure 4.6 Insect-protected Bt cotton. Contrast of yield after heavy lepidopteran pest
pressure. Rows on the left are plants containing events 281-24-236 and 3006-210-23 (i.e.
WideStrike) showing a large number of open bolls and harvestable lint. Rows on the
right are plants of the same genetic background as plants on the left but without the Bt
events. Photo credit: Eileen Crosby and Jim Steadman, Bader Rutter.
7.3.4 Bt potato
Bt potato expressing Cry3Aa was developed by Monsanto and commercial-
ized by Monsanto subsidiary NatureMark in 1995. Multiple events were
deregulated and commercialized as NewLeaf potatoes (expressing
Cry3Aa), NewLeaf Plus potatoes (expressing Cry3Aa and the full length
replicase gene of potato leafroll virus) conferring high levels of control of
the Colorado potato beetle and potato leafroll virus (Lawson et al., 2001)
and NewLeaf Y potatoes (expressing Cry3Aa and potato potyvirus
Y coat protein) (Table 4.9).
Cry3Aa-expressing potatoes provided control of L. decemlineata that was
superior to the commercial insecticides available during the time this
Table 4.7 Cotton events approved for cultivation and commercialized
Year approved Non-IR
Developer Event name OECD unique identifier Bt protein(s) Pest spectrum (cultivationUSA) genesa
Monsanto MON531 MON-000531-6 Cry1Ac Lepidoptera 1995 nptII
aad
Calgene Inc. 31807 Cry1Ac Lepidoptera 1998 bxn
31808 nptII
Dow AgroSciences 281-24-236b DAS-24236-5 Cry1Fa Lepidoptera 2004 pat
b
Dow AgroSciences 3006-210-23 DAS-21023-5 Cry1Ac Lepidoptera 2004 pat
c
Monsanto MON15985 MON-15985-7 Cry1Ac Lepidoptera 2005 nptII
Cry2Ab aad
uidA
Bayer CropSciences GHB119b BCS-GH005-8 Cry2Ae Lepidoptera 2011 bar
b
Bayer CropSciences T304-40 BCS-GH004-7 Cry1Ab Lepidoptera 2011 bar
b
Syngenta COT 102 SYN-IR102-7 Vip3Aa Lepidoptera 2011 aph4
a
Non-IR genes nptII: a selectable marker which confers the ability to metabolize the antibiotics neomycin and kanamycin in plant tissue. aad: a selectable marker used in
the creation of the gene construct and is not expressed in plants. bxn: confers tolerance to the herbicide bromoxinil in plants. pat: a selectable marker which confers
tolerance to the herbicide glufosinate ammonium in plant tissue. uidA(GUS): a scorable marker enabling visual identification of transformed plants. bar: a selectable
marker which confers tolerance to the herbicide glufosinate ammonium in plant tissue. aph4: a selectable marker which confers the ability to metabolize the antibiotic
hygromycin in plant tissue.
b
Commercialized only as a breeding stack.
c
Created via biolistic transformation of germplasm containing Event MON531 (Cry1Ac) with Cry2Ab2.
Table 4.8 Bt cotton products currently sold in the United States
a
Herbicide tolerance events MON88913: glyphosate tolerance trait. LLCotton25: glufosinate ammonium tolerance trait. GBH614: glyphosate tolerance trait.
Shaded boxes indicate that the event is present in the product. See Table 4.7 for Bt genes and pest spectrum associated with Bt events.
Table 4.9 Soybean and potato events approved for cultivation and commercialized
OECD unique Year approved Non-IR
Crop Developer Event name identifier Bt protein(s) Pest spectrum (cultivationUSA) genesa
Potato Monsanto BT6 NMK-89812-3 Cry3A Coleoptera 1995 nptII
BT10 NMK-89175-5
BT12 NMK-89601-8
BT16 NMK-89167-6
BT17 NMK-89593-9
BT18 NMK-89906-7
BT23 NMK-89675-1
Potato Monsanto ATBT04-6 NMK-89761-6 Cry3A Coleoptera 1996 nptII
ATBT04-30 NMK-89613-2
ATBT04-36 NMK-89279-1
SPBT02-5 NMK-89576-1
Potato Monsanto RBMT21-129 NMK-89684-1 Cry3A Coleoptera 1998 nptII
RBMT21-350 NMK-89185-6 aad
RBMT22-082 NMK-89896-6 plrv orf1
plrv orf2
Potato Monsanto RBMT15-101 NMK-89653-6 Cry3A Coleoptera 1999 nptII
SEMT15-02 NMK-89935-9 aad
SEMT15-15 NMK-89930-4 pvy CP
Soybean Monsanto MON87701b MON-87701-2 Cry1Ac Lepidoptera 2011
a
Non-IR genes: nptII: a selectable marker which confers the abilty to metabolize the antibiotics neomycin and kanamycin in plant tissue. aad: a selectable marker used in
the creation of the gene construct and is not expressed in plants. plrv orf1: a viral protein which confers resistance to the potato leafroll virus. plrv orf2: a viral protein which
confers resistance to the potato leafroll virus. pvy CP: a viral coat protein which confers resistance to potato potyvirus Y.
b
Commercialized only as a breeding stack.
230 Kenneth E. Narva et al.
2007; Tabashnik et al., 2011), though this technology has yet to be devel-
oped for commercial application.
9. CONCLUSIONS
It is safe to say that the introduction of Bt crops revolutionized agri-
cultural pest control. Success in the discovery of Bt genes for use in crops
leveraged a long history of development of Bt as a biopesticide. Application
of all of the tools of modern biotechnology was required to genetically engi-
neer crops capable of producing these proteins, and new approaches in many
fields, including plant breeding and regulatory science, were developed to
bring products to the market. The result is crops with a previously unknown
capacity to resist pest damage leading to significant economic, environmen-
tal and societal benefits ( James, 2013).
Nearly 20 years after the introduction of the first Bt crops, the future of
this technology remains bright. Insuring that farmers and society can con-
tinue to reap its benefits remains a priority for technology providers as is
evidenced by the development of pyramided products designed to slow
the inevitable evolutionary response of pest populations to the selection
pressure imposed by this highly effective control. Evidence of the
on-going commitment to commercializing mechanism of action pyramids
is seen in recently announced products in the late stages of development
including: DuPont Pioneers corn event DP4114, a molecular stack of
the Bt proteins contained in Herculex XTRA, which will only be com-
mercialized as a pyramid with other modes of action; Monsantos corn event
MON87411, a pyramid of Cry3Bb and a novel RNAi mechanism of
action, that will be commercialized as a pyramid with Dow AgroSciences
Bt Crops 233
ACKNOWLEDGEMENTS
We would like to thank E. Bouquet, M. Garcia, T. Hey, S. Jayne, R. Maciak and A. Mel for
their review and critique of earlier versions of this chapter. We would also like to thank the
editors of this volume for their constructive comments and patience.
REFERENCES
Abbas, H.K., Zablotowicz, R.M., Weaver, M.A., Shier, W.T., Bruns, H.A., Bellaloui, N.,
Accinelli, C., Abel, C.A., 2013. Implications of Bt traits on mycotoxin contamination in
maize: overview and recent experimental results in southern United States. J. Agric.
Food Chem. 61, 1175911770.
Adang, M.J., Staver, M.J., Rocheleau, T.A., Leighton, J., Barker, R.F., Thompson, D.V.,
1985. Characterized full-length and truncated plasmid clones of the crystal protein of
Bacillus thuringiensis subsp. kurstaki HD-73 and their toxicity to Manduca sexta. Gene
36, 289300.
Adang, M., Brody, M., Cardineau, G., Eagan, N., Roush, R., Shewmaker, C., Jones, A.,
Oakes, J., McBride, K., 1993. The reconstruction and expression of a Bacillus thuringiensis
cryIIIA gene in protoplasts and potato plants. Plant Mol. Biol. 21, 11311145.
Alstad, D.N., Andow, D.A., 1995. Managing the evolution of insect resistance to transgenic
plants. Science 268, 18941896.
Angus, T.A., 1954. A bacterial toxin paralysing silkworm larvae. Nature 173, 545546.
Arantes, O., Lereclus, D., 1991. Construction of cloning vectors for Bacillus thuringiensis.
Gene 108, 115119.
Archer, T.L., Patrick, C., Schuster, G., Cronholm, G., Bynum Jr., E.D., Morrison, W.P.,
2001. Ear and shank damage by corn borers and corn earworms to four events of Bacillus
thuriengiensis transgenic maize. Crop Prot. 20, 139144.
Arrieta, G., Espinoza, A.M., 2006. Characterization of a Bacillus thuringiensis strain collection
isolated from diverse Costa Rican natural ecosystems. Rev. Biol. Trop. 54, 1327.
Barfoot, P., Brookes, G., 2014. Key global environmental impacts of genetically modified
(GM) crop use 19962012. GM Crops Food Biotechnol. Agric. Food Chain 5, 1021.
Barloy, F., Delecluse, A., Nicolas, L., Lecadet, M.M., 1996. Cloning and expression of the
first anaerobic toxin gene from Clostridium bifermentans subsp. malaysia, encoding a new
mosquitocidal protein with homologies to Bacillus thuringiensis delta-endotoxins.
J. Bacteriol. 178, 30993105.
Baum, J.A., Johnson, T.B., Carlton, B.C., 1998. Bacillus thuringiensis: natural and recombi-
nant bioinsecticide products. In: Hall, F.R., Menn, J.J. (Eds.), Biopesticides: Use and
Delivery, pp. 189209.
234 Kenneth E. Narva et al.
Baxter, S.W., Badenes-Perez, F.R., Morrison, A., Vogel, H., Crickmore, N., Kain, W.,
Wang, P., Heckel, D.G., Jiggins, C.D., 2011. Parallel evolution of Bacillus thuringiensis
toxin resistance in lepidoptera. Genetics 189, 675679.
Ben-Dov, E., 2014. Bacillus thuringiensis subsp. israelensis and its Dipteran-specific toxins.
Toxins 6, 12221243.
Bergamasco, V.B., Mendes, D.R., Fernandes, O.A., Desiderio, J.A., Lemos, M.V., 2013.
Bacillus thuringiensis Cry1Ia10 and Vip3Aa protein interactions and their toxicity in
Spodoptera spp. (Lepidoptera). J. Invertebr. Pathol. 112, 152158.
Berliner, E., 1915. Ueber die schlaVsucht der Ephestia kuhniella und Bac. thuringiensis n. sp.
Z. Angew. Entomol. 2, 2156.
Bernardi, O., Sorgatto, R.J., Barbosa, A.D., Domingues, F.A., Dourado, P.M., Carvalho, R.A.,
Martinelli, S., Head, G.P., Omoto, C., 2014. Low susceptibility of Spodoptera cosmioides,
Spodoptera eridania and Spodoptera frugiperda (Lepidoptera: Noctuidae) to genetically-
modified soybean expressing Cry1Ac protein. Crop Prot. 58, 3340.
Berry, C., 2012. The bacterium, Lysinibacillus sphaericus, as an insect pathogen. J. Invertebr.
Pathol. 109, 110.
Betz, F.S., Hammond, B.G., Fuchs, R.L., 2000. Safety and advantages of Bacillus thuringiensis-
protected plants to control insect pests. Regul. Toxicol. Pharmacol. 32, 156173.
Blanco, C.A., 2012. Heliothis virescens and Bt cotton in the United States. GM Crops Food
Biotechnol. Agric. Food Chain 3, 201212.
Boiteau, G., Osborn, W.P.L., Drew, M.E., 1997. Residual activity of lmidacloprid control-
ling colorado potato beetle (Coleoptera: Chrysomelidae) and three species of potato col-
onizing aphids (Homoptera: Aphidae). J. Econ. Entomol. 90, 309319.
Bolivar, F., Rodriguez, R.L., Betlach, M.C., Boyer, H.W., 1977. Construction and charac-
terization of new cloning vehicles I. Ampicillin-resistant derivatives of the plasmid
pMB9. Gene 2, 7593.
Bommireddy, P.L., Leonard, B.R., Temple, J., Price, P., Emfinger, K., Cook, D.,
Hardke, J.T., 2011. Field performance and seasonal efficacy profiles of transgenic cotton
lines expressing Vip3A and VipCot against Helicoverpa zea (Boddie) and heliothis vir-
escens (F.). J. Cotton Sci. 15, 251259.
Boonserm, P., Davis, P., Ellar, D.J., Li, J., 2005. Crystal structure of the mosquito-larvicidal
toxin Cry4Ba and its biological implications. J. Mol. Biol. 348, 363382.
Boonserm, P., Mo, M., Angsuthanasombat, C., Lescar, J., 2006. Structure of the functional
form of the mosquito larvicidal Cry4Aa toxin from Bacillus thuringiensis at a
2.8-angstrom resolution. J. Bacteriol. 188, 33913401.
Bravo, A., Sarabia, S., Lopez, L., Ontiveros, H., Abarca, C., Ortiz, A., Ortiz, M., Lina, L.,
Villalobos, F.J., Pena, G., Nunez-Valdez, M.E., Soberon, M., Quintero, R., 1998.
Characterization of cry genes in a Mexican Bacillus thuringiensis strain collection. Appl.
Environ. Microbiol. 64, 49654972.
Brookes, G., Barfoot, P., 2014. Economic impact of GM crops: the global income and
production effects 19962012. GM Crops Food Biotechnol. Agric. Food Chain
5, 3747.
Buntin, G.D., 2008. Corn expressing Cry1Ab or Cry1F endotoxin for fall armyworm and
corn earworm (Lepidoptera: Noctuidae) management in field corn for grain production.
Fla. Entomol. 91, 523530.
Buntin, G.D., All, J.N., Lee, R.D., Wilson, D.M., 2004. Plant-incorporated Bacillus thur-
ingiensis resistance for control of fall armyworm and corn earworm (Lepidoptera:
Noctuidae) in corn. J. Econ. Entomol. 97, 16031611.
Burkness, E.C., Dively, G., Patton, T., Morey, A.C., Hutchison, W.D., 2010. Novel Vip3A
Bacillus thuringiensis (Bt) maize approaches high-dose efficacy against Helicoverpa zea
(Lepidoptera: Noctuidae) under field conditions: implications for resistance manage-
ment. GM Crops 1, 337343.
Bt Crops 235
Butko, P., 2003. Cytolytic toxin Cyt1A and its mechanism of membrane damage: data and
hypotheses. Appl. Environ. Microbiol. 69, 24152422.
Caccia, S., Hernandez-Rodriguez, C.S., Mahon, R.J., Downes, S., James, W.,
Bautsoens, N., Van Rie, J., Ferre, J., 2010. Binding site alteration is responsible for
field-isolated resistance to Bacillus thuringiensis Cry2A insecticidal proteins in two Hel-
icoverpa species. PLoS One 5, e9975.
Caramori, T., Albertini, A.M., Galizzi, A., 1991. In vivo generation of hybrids between two
Bacillus thuringiensis insect-toxin-encoding genes. Gene 98, 3744.
Carlton, B.C., Gawron-Burke, C., 1993. Genetic Improvement of Bacillus thuringiensis for
Bioinsecticide Development. Marcel Dekker Inc., New York, NY.
Carozzi, N.B., Kramer, V.C., Warren, G.W., Evola, S., Koziel, M.G., 1991. Prediction of
insecticidal activity of Bacillus thuringiensis strains by polymerase chain reaction product
profiles. Appl. Environ. Microbiol. 57, 30573061.
Carroll, J., Li, J., Ellar, D.J., 1989. Proteolytic processing of a coleopteran-specific
delta-endotoxin produced by Bacillus thuringiensis var. tenebrionis. Biochem. J.
261, 99105.
Carroll, J., Convents, D., Van Damme, J., Boets, A., Van Rie, J., Ellar, D.J., 1997. Intramo-
lecular proteolytic cleavage of Bacillus thuringiensis Cry3A delta-endotoxin may facilitate
its coleopteran toxicity. J. Invertebr. Pathol. 70, 4149.
Carroll, M.W., Head, G., Caprio, M.A., 2012. When and where a seed mix refuge makes
sense for managing insect resistance to Bt plants. Crop Prot. 38, 7479.
Catangui, M.A., 2003. Transgenic Bacillus thuringiensis corn hybrid performance against
univoltine ecotype European corn borer (Lepidoptera: Crambidae) in South Dakota.
J. Econ. Entomol. 96, 957968.
CERA, 2012. GM Crop Database. Center for Environmental Risk Assessment (CERA),
ILSI Research Foundation, Washington D.C.http://cera-gmc.org/index.php?
actiongm_crop_database.
Chang, C., Yu, Y.M., Dai, S.M., Law, S.K., Gill, S.S., 1993. High-level cryIVD and cytA
gene expression in Bacillus thuringiensis does not require the 20-kilodalton protein, and
the coexpressed gene products are synergistic in their toxicity to mosquitoes. Appl. Envi-
ron. Microbiol. 59, 815821.
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, 84658470.
Cohen, S., Dym, O., Albeck, S., Ben-Dov, E., Cahan, R., Firer, M., Zaritsky, A., 2008.
High-resolution crystal structure of activated Cyt2Ba monomer from Bacillus thur-
ingiensis subsp. israelensis. J. Mol. Biol. 380, 820827.
Cohen, S., Albeck, S., Ben-Dov, E., Cahan, R., Firer, M., Zaritsky, A., Dym, O., 2011.
Cyt1Aa toxin: crystal structure reveals implications for its membrane-perforating func-
tion. J. Mol. Biol. 413, 804814.
Contreras, E., Schoppmeier, M., Real, M.D., Rausell, C., 2013. Sodium solute symporter
and cadherin proteins act as Bacillus thuringiensis Cry3Ba toxin functional receptors in Tri-
bolium castaneum. J. Biol. Chem. 288, 1801318021.
Corbin, D.R., Romano, C.P., 2006. Plants transformed to express Cry2A -endotoxins.
U.S. patent number 7,064,249.
Craveiro, K.I., Gomes Junior, J.E., Silva, M.C., Macedo, L.L., Lucena, W.A., Silva, M.S., de
Souza Junior, J.D., Oliveira, G.R., de Magalhaes, M.T., Santiago, A.D., Grossi-de-Sa,
M.F., 2010. Variant Cry1Ia toxins generated by DNA shuffling are active against sugar-
cane giant borer. J. Biotechnol. 145, 215221.
Crickmore, N., Zeigler, D.R., Feitelson, J., Schnepf, E., Van Rie, J., Lereclus, D., Baum, J.,
Dean, D.H., 1998. Revision of the nomenclature for the Bacillus thuringiensis pesticidal
crystal proteins. Microbiol. Mol. Biol. Rev. 62, 807813.
236 Kenneth E. Narva et al.
Crickmore, N., Baum, J., Bravo, A., Lereclus, D., Narva, K., Sampson, K., Schnepf, E.,
Sun, M., Zeigler, D.R., 2014. Bacillus thuringiensis toxin nomenclature. http://
www.btnomenclature.info/.
Dean, D.H., Rajamohan, F., Lee, M.K., Wu, S.J., Chen, X.J., Alcantara, E., Hussain, S.R.,
1996. Probing the mechanism of action of Bacillus thuringiensis insecticidal proteins by
site-directed mutagenesisa minireview. Gene 179, 111117.
de Barjac, H., Bonnefoi, A., 1962. Essai de classification biochemique et serologique de 24
sourches de Bacillus du type B. thuringiensis. Entomophaga 7, 531.
de Barjac, H., Bonnefoi, A., 1968. A classification of strains of Bacillus thuringiensis Berliner
with a key to their differentiation. J. Invertebr. Pathol. 11, 335347.
de Freitas Bueno, R.C.O., de Freitas Bueno, A., Moscardi, F., Postali Parra, J.R., Hoffmann-
Campo, C.B., 2011. Lepidopteran larva consumption of soybean foliage: basis for devel-
oping multiple-species economic thresholds for pest management decisions. Pest Manag.
Sci. 67, 170174.
de Maagd, R.A., Kwa, M.S., van der Klei, H., Yamamoto, T., Schipper, B., Vlak, J.M.,
Stiekema, W.J., Bosch, D., 1996a. Domain III substitution in Bacillus thuringiensis
delta-endotoxin CryIA(b) results in superior toxicity for Spodoptera exigua and altered
membrane protein recognition. Appl. Environ. Microbiol. 62, 15371543.
de Maagd, R.A., van der Klei, H., Bakker, P.L., Stiekema, W.J., Bosch, D., 1996b. Different
domains of Bacillus thuringiensis delta-endotoxins can bind to insect midgut membrane
proteins on ligand blots. Appl. Environ. Microbiol. 62, 27532757.
de Maagd, R.A., Bravo, A., Crickmore, N., 2001. How Bacillus thuringiensis has evolved spe-
cific toxins to colonize the insect world. Trends Genet. 17, 193199.
de Maagd, R.A., Bravo, A., Berry, C., Crickmore, N., Schnepf, H.E., 2003. Structure,
diversity, and evolution of protein toxins from spore-forming entomopathogenic bacte-
ria. Annu. Rev. Genet. 37, 409433.
DiFonzo, C., Cullen, E.M., 2013. Handy Bt Trait Table. University of Wisconsin. http://
labs.russell.wisc.edu/cullenlab/extension/extension-publications/.
DiLeo, M.V., Bakker, M.d., Chu, E.Y., Hoekenga, O.A., 2014. An assessment of the relative
influences of genetic background, functional diversity at major regulatory genes, and
transgenic constructs on the tomato fruit metabolome. Plant Genome 7. http://dx.
doi.org/10.3835/plantgenome2013.06.0021.
Dimock, M., Turner, J., Lampel, J., 1993. Endophytic microorganisms for delivery of genet-
ically engineered microbial pesticides in plants. In: Kim, L. (Ed.), Advanced Engineered
Biopesticides. Marcel Dekker Inc., New York, NY
Donovan, W.P., Rupar, M.J., Slaney, A.C., Malvar, T., Gawron-Burke, M.C.,
Johnson, T.B., 1992. Characterization of two genes encoding Bacillus thuringiensis insec-
ticidal crystal proteins toxic to Coleoptera species. Appl. Environ. Microbiol.
58, 39213927.
Downes, S., Mahon, R., 2012. Successes and challenges of managing resistance in Helicoverpa
armigera to Bt cotton in Australia. GM Crops Food 3, 228234.
Dulmage, H.D., 1981. Insecticidal activity of isolates of Bacillus thuringiensis and their poten-
tial for pest control. In: Burges, H.D. (Ed.), Microbial Control of Pests and Diseases,
1970-1980. Academic Press, London, pp. 193222.
Eichenseer, H., Strohbehn, R., Burks, J., 2008. Frequency and severity of western bean cut-
worm (Lepidoptera: Noctuidae) ear damage in transgenic corn hybrids expressing differ-
ent Bacillus thuringiensis cry toxins. J. Econ. Entomol. 101, 555563.
Ellis, R.T., Stockhoff, B.A., Stamp, L., Schnepf, H.E., Schwab, G.E., Knuth, M., Russell, J.,
Cardineau, G.A., Narva, K.E., 2002. Novel Bacillus thuringiensis binary insecticidal crystal
proteins active on western corn rootworm, Diabrotica virgifera virgifera LeConte. Appl.
Environ. Microbiol. 68, 11371145.
Bt Crops 237
Estruch, J.J., Warren, G.W., Mullins, M.A., Nye, G.J., Craig, J.A., Koziel, M.G., 1996.
Vip3A, a novel Bacillus thuringiensis vegetative insecticidal protein with a wide spectrum
of activities against lepidopteran insects. Proc. Natl. Acad. Sci. U. S. A. 93, 53895394.
Fang, J., Xu, X., Wang, P., Zhao, J.Z., Shelton, A.M., Cheng, J., Feng, M.G., Shen, Z.,
2007. Characterization of chimeric Bacillus thuringiensis Vip3 toxins. Appl. Environ.
Microbiol. 73, 956961.
Fang, Y., Li, Z., Liu, J., Shu, C., Wang, X., Zhang, X., Yu, X., Zhao, D., Liu, G., Hu, S.,
Zhang, J., Al-Mssallem, I., Yu, J., 2011. A pangenomic study of Bacillus thuringiensis.
J. Genet. Genomics 38, 567576.
FAO Expert Consultation on Genetically modified organisms in crop production and
their effects on the environment: methodologies for monitoring and the way ahead.
2005. Food and Agriculture Organization of the United Nations. Rome, Italy.
http://www.fao.org/docrep/008/ae738e/ae738e00.htm TopOfPagehttp://
www.fao.org/docrep/008/ae738e/ae738e00.htm#TopOfPage.
Farinos, G.P., De La Poza, M., Hernandez-Crespo, P., Ortego, F., Castanera, P., 2004.
Resistance monitoring of field populations of the corn borers Sesamia nonagrioides
and Ostrinia nubilalis after 5 years of Bt maize cultivation in Spain. Entomol. Exp. Appl.
110, 2330.
Farinos, G.P., De la Poza, M., Ortego, F., Castanera, P., 2012. Susceptibility to the Cry1F
toxin of field populations of Sesamia nonagrioides (Lepidoptera: Noctuidae) in Mediter-
ranean maize cultivation regions. J. Econ. Entomol. 105, 214221.
Federici, B.A., 2005. Insecticidal bacteria: an overwhelming success for invertebrate pathol-
ogy. J. Invertebr. Pathol. 89, 3038.
Federici, B.A., Bauer, L.S., 1998. Cyt1Aa protein of Bacillus thuringiensis is toxic to the cot-
tonwood leaf beetle, chrysomela scripta, and suppresses high levels of resistance to
Cry3Aa. Appl. Environ. Microbiol. 64, 43684371.
Fedoroff, N.V., 2011. Burdensome and unnecessary regulation. GM Crops 2, 8788.
Feitelson, J.S., Payne, J., Kim, L., 1992. Bacillus thuringiensis: insects and beyond. Nat. Bio-
technol. 10, 271275.
Finer, J., Dhillon, T., 2007. Transgenic plant production. In: Stewart, C.N. (Ed.), Plant Bio-
technology and Genetics: Principles, Techniques, and Applications. John Wiley & Sons,
Inc., Hoboken, NJ, pp. 245274
Fischhoff, D.A., Bowdish, K.S., Perlak, F.J., Marrone, P.G., McCormick, S.M.,
Niedermeyer, J.G., Dean, D.A., Kusano-Kretzmer, K., Mayer, E.J., Rochester, D.E.,
Rogers, S.G., Fraley, R.T., 1987. Insect tolerant transgenic tomato plants. Nat. Bio-
technol. 5, 807813.
Fox, J.L., 2001. EPA re-evaluates StarLink license. Nat. Biotechnol. 19, 11.
Gaertner, F.H., Quick, T.C., Thompson, M.A., 1993. Cell cap: an encapsulation system for
insecticidal biotoxin proteins. In: Kim, L. (Ed.), Advanced Engineered Biopesticides.
Marcel Dekker Inc., New York, NY, pp. 7384.
Gahan, L.J., Gould, F., Heckel, D.G., 2001. Identification of a gene associated with Bt resis-
tance in Heliothis virescens. Science 293, 857860.
Gahan, L.J., Pauchet, Y., Vogel, H., Heckel, D.G., 2010. An ABC transporter mutation is cor-
related with insect resistance to Bacillus thuringiensis Cry1Ac toxin. PLoS Genet. 6, e1001248.
Galitsky, N., Cody, V., Wojtczak, A., Ghosh, D., Luft, J.R., Pangborn, W., English, L.,
2001. Structure of the insecticidal bacterial delta-endotoxin Cry3Bb1 of Bacillus thur-
ingiensis. Acta Crystallogr. D Biol. Crystallogr. 57, 11011109.
Gassmann, A.J., 2012. Field-evolved resistance to Bt maize by western corn rootworm: pre-
dictions from the laboratory and effects in the field. J. Invertebr. Pathol. 110, 287293.
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.
238 Kenneth E. Narva et al.
Gassmann, A.J., Petzold-Maxwell, J.L., Clifton, E.H., Dunbar, M.W., Hoffmann, A.M.,
Ingber, D.A., Keweshan, R.S., 2014. Field-evolved resistance by western corn
rootworm to multiple Bacillus thuringiensis toxins in transgenic maize. Proc. Natl. Acad.
Sci. 111, 51415146.
Ghimire, M.N., Huang, F., Leonard, R., Head, G.P., Yang, Y., 2011. Susceptibility of
Cry1Ab-susceptible and -resistant sugarcane borer to transgenic corn plants containing
single or pyramided Bacillus thuringiensis genes. Crop Prot. 30, 7481.
Goldberg, L.J., Margalit, J., 1977. A bacterial spore demonstrating rapid larvicidal activity
against Anopheles sergentii, Uranotaeniaunguiculata, Culex univittatus, Aedes aegypti, and
Culex pipiens. Mosq. News 37, 317324.
Gomez, I., Pardo-Lopez, L., Munoz-Garay, C., Fernandez, L.E., Perez, C., Sanchez, J.,
Soberon, M., Bravo, A., 2007. Role of receptor interaction in the mode of action of
insecticidal Cry and Cyt toxins produced by Bacillus thuringiensis. Peptides 28, 169173.
Gouffon, C., Van Vliet, A., Van Rie, J., Jansens, S., Jurat-Fuentes, J.L., 2011. Binding sites
for Bacillus thuringiensis Cry2Ae toxin on heliothine brush border membrane vesicles
are not shared with Cry1A, Cry1F, or Vip3A toxin. Appl. Environ. Microbiol.
77, 31823188.
Gould, F., 1998. Sustainability of transgenic insecticidal cultivars: integrating pest genetics
and ecology. Annu. Rev. Entomol. 43, 701726.
Gould, F., Blair, N., Reid, M., Rennie, T.L., Lopez, J., Micinski, S., 2002. Bacillus thur-
ingiensis-toxin resistance management: stable isotope assessment of alternate host use
by Helicoverpa zea. Proc. Natl. Acad. Sci. U. S. A. 99, 1658116586.
Griffitts, J.S., Haslam, S.M., Yang, T., Garczynski, S.F., Mulloy, B., Morris, H.,
Cremer, P.S., Dell, A., Adang, M.J., Aroian, R.V., 2005. Glycolipids as receptors for
Bacillus thuringiensis crystal toxin. Science 307, 922925.
Grochulski, P., Masson, L., Borisova, S., Pusztai-Carey, M., Schwartz, J.L., Brousseau, R.,
Cygler, M., 1995. Bacillus thuringiensis CryIA(a) insecticidal toxin: crystal structure and
channel formation. J. Mol. Biol. 254, 447464.
Gunning, R.V., Dang, H.T., Kemp, F.C., Nicholson, I.C., Moores, G.D., 2005. New resis-
tance mechanism in Helicoverpa armigera threatens transgenic crops expressing Bacillus
thuringiensis Cry1Ac toxin. Appl. Environ. Microbiol. 71, 25582563.
Guo, S., Ye, S., Liu, Y., Wei, L., Xue, J., Wu, H., Song, F., Zhang, J., Wu, X., Huang, D.,
Rao, Z., 2009. Crystal structure of Bacillus thuringiensis Cry8Ea1: an insecticidal toxin
toxic to underground pests, the larvae of Holotrichia parallela. J. Struct. Biol.
168, 259266.
Gustafson, D.I., Head, G.P., Caprio, M.A., 2006. Modeling the impact of alternative hosts
on Helicoverpa zea adaptation to bollgard cotton. J. Econ. Entomol. 99, 21162124.
Hammond, B.G., Jez, J.M., 2011. Impact of food processing on the dietary risk assessment for
proteins introduced into biotechnology-derived soybean and corn crops. Food Chem.
Toxicol. 49, 711721.
Hammond, R.B., Higley, L.G., Pedigo, L.P., Bledsoe, L., Spomer, S.M., DeGooyer, T.A.,
2000. Simulated Insect Defoliation on Soybean: Influence of Row Width. J. Econ.
Entomol. 93, 14291436.
Han, S., Craig, J.A., Putnam, C.D., Carozzi, N.B., Tainer, J.A., 1999. Evolution and mech-
anism from structures of an ADP-ribosylating toxin and NAD complex. Nat. Struct.
Mol. Biol. 6, 932936.
Hannay, C.L., 1953. Crystalline inclusions in aerobic sporeforming bacteria. Nature 172, 1004.
Hansen, G., Wright, M.S., 1999. Recent advances in the transformation of plants. Trends
Plant Sci. 4, 226231.
Hardke, J.T., Leonard, B.R., Huang, F., Jackson, R.E., 2011. Damage and survivorship of
fall armyworm (Lepidoptera: Noctuidae) on transgenic field corn expressing Bacillus
thuringiensis Cry proteins. Crop Prot. 30, 168172.
Bt Crops 239
He, K., Wang, Z., Zhou, D., Wen, L., Song, Y., Yao, Z., 2003. Evaluation of transgenic Bt
corn for resistance to the Asian corn borer (Lepidoptera: Pyralidae). J. Econ. Entomol.
96, 935940.
Head, G.P., Greenplate, J., 2012. The design and implementation of insect resistance man-
agement programs for Bt crops. GM Crops Food 3, 144153.
Head, G., Carroll, M., Clark, T., Galvan, T., Huckaba, R.M., Price, P., Samuel, L.,
Storer, N.P., 2014. Efficacy of SmartStax(r) insect-protected corn hybrids against corn
rootworm: the value of pyramiding the Cry3Bb1 and Cry34/35Ab1 proteins. Crop Prot.
57, 3847.
Heckel, D.G., Gahan, L.J., Baxter, S.W., Zhao, J.Z., Shelton, A.M., Gould, F.,
Tabashnik, B.E., 2007. The diversity of Bt resistance genes in species of Lepidoptera.
J. Invertebr. Pathol. 95, 192197.
Hellmich, R.L., Siegfried, B.D., Sears, M.K., Stanley-Horn, D.E., Daniels, M.J.,
Mattila, H.R., Spencer, T., Bidne, K.G., Lewis, L.C., 2001. Monarch larvae sensitivity
to Bacillus thuringiensis- purified proteins and pollen. Proc. Natl. Acad. Sci. U. S. A.
98, 1192511930.
Herman, R.A., Price, W.D., 2013. Unintended compositional changes in genetically mod-
ified (gm) crops: 20 years of research. J. Agric. Food Chem. 61, 1169511701.
Herman, R.A., Scherer, P.N., Young, D.L., Mihaliak, C.A., Meade, T., Woodsworth, A.T.,
Stockhoff, B.A., Narva, K.E., 2002. Binary insecticidal crystal protein from Bacillus thur-
ingiensis, strain PS149B1: effects of individual protein components and mixtures in lab-
oratory bioassays. J. Econ. Entomol. 95, 635639.
Herman, R.A., Chassy, B.M., Parrott, W., 2009. Compositional assessment of transgenic
crops: an idea whose time has passed. Trends Biotechnol. 27, 555557.
Hernandez-Martinez, P., Naseri, B., Navarro-Cerrillo, G., Escriche, B., Ferre, J., Herrero, S.,
2010. Increase in midgut microbiota load induces an apparent immune priming and
increases tolerance to Bacillus thuringiensis. Environ. Microbiol. 12, 27302737.
Hernandez-Martinez, P., Hernandez-Rodriguez, C.S., Rie, J.V., Escriche, B., Ferre, J.,
2013. Insecticidal activity of Vip3Aa, Vip3Ad, Vip3Ae, and Vip3Af from Bacillus thur-
ingiensis against lepidopteran corn pests. J. Invertebr. Pathol. 113, 7881.
Hernandez-Rodriguez, C.S., Van Vliet, A., Bautsoens, N., Van Rie, J., Ferre, J., 2008. Spe-
cific binding of Bacillus thuringiensis Cry2A insecticidal proteins to a common site in the
midgut of Helicoverpa species. Appl. Environ. Microbiol. 74, 76547659.
Herrero, S., Gechev, T., Bakker, P.L., Moar, W.J., de Maagd, R.A., 2005. Bacillus thur-
ingiensis Cry1Ca-resistant Spodoptera exigua lacks expression of one of four aminopep-
tidase N genes. BMC Genomics 6, 96.
Herrnstadt, C., Gilroy, T.E., Sobieski, D.A., Bennett, B.D., Gaertner, F.H., 1987. Nucle-
otide sequence and deduced amino acid sequence of a coleopteran-active delta-
endotoxin gene from Bacillus thuringiensis subsp. san diego. Gene 57, 3746.
Hofte, H., Whiteley, H.R., 1989. Insecticidal crystal proteins of Bacillus thuringiensis.
Microbiol. Rev. 53, 242255.
Hofte, H., Van Rie, J., Jansens, S., Van Houtven, A., Vanderbruggen, H., Vaeck, M., 1988.
Monoclonal antibody analysis and insecticidal spectrum of three types of lepidopteran-
specific insecticidal crystal proteins of Bacillus thuringiensis. Appl. Environ. Microbiol.
54, 20102017.
Hui, F., Scheib, U., Hu, Y., Sommer, R.J., Aroian, R.V., Ghosh, P., 2012. Structure and
glycolipid binding properties of the nematicidal protein Cry5B. Biochemistry
51, 99119921.
Husz, B., 1930. Field experiments on the application of Bacillus thuringiensis against the corn
borer. Int. Corn Borer Invest. Sci. Rep. 3, 9198.
Ishiwata, S., 1901. On a type of severe Xacherie (sotto disease). Dainihon Sanshi Kaiho
114, 15.
240 Kenneth E. Narva et al.
Ives, A.R., Glaum, P.R., Ziebarth, N.L., Andow, D.A., 2011. The evolution of resistance to
two-toxin pyramid transgenic crops. Ecol. Appl. 21, 503515.
Jackson, R.E., Bradley Jr., J.R., Van Duyn, J.W., Gould, F., 2004. Comparative production
of Helicoverpa zea (Lepidoptera: Noctuidae) from transgenic cotton expressing either one
or two Bacillus thuringiensis proteins with and without insecticide oversprays. J. Econ.
Entomol. 97, 17191725.
James, C., 2013. Global Status of Commercialize/Biotech Crops: 2013. ISAAA, Ithaca, NY,
ISAAA Brief No. 46.
Jones, G.W., Nielsen-Leroux, C., Yang, Y., Yuan, Z., Dumas, V.F., Monnerat, R.G.,
Berry, C., 2007. A new Cry toxin with a unique two-component dependency from
Bacillus sphaericus. FASEB J. 21, 41124120.
Jung, Y.C., Kim, S.U., Cote, J.C., Lecadet, M.M., Chung, Y.S., Bok, S.H., 1998. Charac-
terization of a new Bacillus thuringiensis Subsp. higo strain isolated from rice bran in Korea.
J. Invertebr. Pathol. 71, 9596.
Jurat-Fuentes, J.L., Gahan, L.J., Gould, F.L., Heckel, D.G., Adang, M.J., 2004. The
HevCaLP protein mediates binding specificity of the Cry1A class of Bacillus thuringiensis
toxins in Heliothis virescens. Biochemistry 43, 1429914305.
Kaur, S., 2000. Molecular approaches towards development of novel Bacillus thuringiensis bio-
pesticides. World J. Microbiol. Biotechnol. 16, 781793.
Keller, M., Sneh, B., Strizhov, N., Prudovsky, E., Regev, A., Koncz, C., Schell, J.,
Zilberstein, A., 1996. Digestion of delta-endotoxin by gut proteases may explain reduced
sensitivity of advanced instar larvae of Spodoptera littoralis to CryIC. Insect Biochem. Mol.
Biol. 26, 365373.
Khajuria, C., Buschman, L.L., Chen, M.S., Siegfried, B.D., Zhu, K.Y., 2011. Identification
of a novel aminopeptidase P-like gene (OnAPP) possibly involved in Bt toxicity and
resistance in a major corn pest (Ostrinia nubilalis). PLoS One 6, e23983.
Koziel, M.G., Beland, G.L., Bowman, C., Carozzi, N.B., Crenshaw, R., Crossland, L.,
Dawson, J., Desai, N., Hill, M., Kadwell, S., Launis, K., Lewis, K., Maddox, D.,
McPherson, K., Meghji, M.R., Merlin, E., Rhodes, R., Warren, G.W., Wright, M.,
Evola, S.V., 1993. Field performance of elite transgenic maize plants expressing an insec-
ticidal protein derived from Bacillus thuringiensis. Nat. Biotechnol. 11, 194200.
Krieg, A., Huger, A.M., Langenbruch, G.A., Schnetter, W., 1983. Bacillus thuringiensis var.
tenebrionis: a new pathotype effective against larvae of Coleoptera. Z. Angew. Entomol.
96, 500508.
Kronstad, J.W., Whiteley, H.R., 1986. Three classes of homologous Bacillus thuringiensis
crystal-protein genes. Gene 43, 2940.
Kronstad, J.W., Schnepf, H.E., Whiteley, H.R., 1983. Diversity of locations for Bacillus thur-
ingiensis crystal protein genes. J. Bacteriol. 154, 419428.
Kullik, S.A., Sears, M.K., Schaafsma, A.W., 2011. Sublethal effects of Cry 1F Bt corn and
clothianidin on black cutworm (Lepidoptera: Noctuidae) larval development. J. Econ.
Entomol. 104, 484493.
Ladics, G.S., Cressman, R.F., Herouet-Guicheney, C., Herman, R.A., Privalle, L., Song, P.,
Ward, J.M., McClain, S., 2011. Bioinformatics and the allergy assessment of agricultural
biotechnology products: industry practices and recommendations. Regul. Toxicol.
Pharmacol. 60, 4653.
Lampel, J.S., Canter, G.L., Dimock, M.B., Kelly, J.L., Anderson, J.J., Uratani, B.B.,
Foulke, J.S., Turner, J.T., 1994. Integrative cloning, expression, and stability of the
cryIA(c) gene from Bacillus thuringiensis subsp. kurstaki in a recombinant strain of
Clavibacter xyli subsp. cynodontis. Appl. Environ. Microbiol. 60, 501508.
Lawson, E.C., Weiss, J.D., Thomas, P.E., Kaniewski, W.K., 2001. NewLeaf Plus(r) Russet
Burbank potatoes: replicase-mediated resistance to potato leafroll virus. Mol. Breed.
7, 112.
Bt Crops 241
Lecadet, M.M., Chaufaux, J., Ribier, J., Lereclus, D., 1992. Construction of novel Bacillus
thuringiensis strains with different insecticidal activities by transduction and transforma-
tion. Appl. Environ. Microbiol. 58, 840849.
Lee, M.K., Walters, F.S., Hart, H., Palekar, N., Chen, J.S., 2003. The mode of action of the
Bacillus thuringiensis vegetative insecticidal protein Vip3A differs from that of Cry1Ab
delta-endotoxin. Appl. Environ. Microbiol. 69, 46484657.
Lee, M.K., Miles, P., Chen, J.S., 2006. Brush border membrane binding properties of Bacillus
thuringiensis Vip3A toxin to Heliothis virescens and Helicoverpa zea midguts. Biochem.
Biophys. Res. Commun. 339, 10431047.
Lee, H., Park, S.Y., Zhang, Z.J., 2013. An overview of genetic transformation of soybean.
In: Board, J. (Ed.), A Comprehensive Survey of International Soybean Research -
Genetics. Physiology, Agronomy and Nitrogen Relationships. InTech, pp. 439506.
http://www.intechopen.com/books/a-comprehensive-survey-of-international-soybean-
research-genetics-physiology-agronomy-and-nitrogen-relationships/an-overview-
of-genetic-transformation-of-soybean.
Letowski, J., Bravo, A., Brousseau, R., Masson, L., 2005. Assessment of cry1 gene contents of
Bacillus thuringiensis strains by use of DNA microarrays. Appl. Environ. Microbiol.
71, 53915398.
Li, J.D., Carroll, J., Ellar, D.J., 1991. Crystal structure of insecticidal delta-endotoxin from
Bacillus thuringiensis at 2.5 A resolution. Nature 353, 815821.
Li, J., Koni, P.A., Ellar, D.J., 1996. Structure of the mosquitocidal delta-endotoxin CytB
from Bacillus thuringiensis sp. kyushuensis and implications for membrane pore formation.
J. Mol. Biol. 257, 129152.
Li, H., Oppert, B., Higgins, R.A., Huang, F., Zhu, K.Y., Buschman, L.L., 2004. Compar-
ative analysis of proteinase activities of Bacillus thuringiensis-resistant and -susceptible
Ostrinia nubilalis (Lepidoptera: Crambidae). Insect Biochem. Mol. Biol. 34, 753762.
Li, H., Olson, M., Lin, G., Hey, T., Tan, S.Y., Narva, K.E., 2013. Bacillus thuringiensis
Cry34Ab1/Cry35Ab1 interactions with western corn rootworm midgut membrane
binding sites. PLoS One 8, e53079.
Likitvivatanavong, S., Chen, J., Bravo, A., Soberon, M., Gill, S.S., 2011a. Cadherin, alkaline
phosphatase, and aminopeptidase N as receptors of Cry11Ba toxin from Bacillus thur-
ingiensis subsp. jegathesan in Aedes aegypti. Appl. Environ. Microbiol. 77, 2431.
Likitvivatanavong, S., Chen, J., Evans, A.M., Bravo, A., Soberon, M., Gill, S.S., 2011b.
Multiple receptors as targets of Cry toxins in mosquitoes. J. Agric. Food Chem.
59, 28292838.
Liu, X.S., Dean, D.H., 2006. Redesigning Bacillus thuringiensis Cry1Aa toxin into a mosquito
toxin. Protein Eng. Des. Sel. 19, 107111.
Luttrell, R.G., Jackson, R.E., 2012. Helicoverpa zea and Bt cotton in the United States. GM
Crops Food Biotechnol. Agric. Food Chain 3, 213227.
Ma, G., Schmidt, O., Keller, M., 2012. Pre-feeding of a glycolipid binding protein LEC-8
from Caenorhabditis elegans revealed enhanced tolerance to Cry1Ac toxin in Helicoverpa
armigera. Results Immunol. 2, 97103.
MacIntosh, S.C., 2010. Managing the risk of insect resistance to transgenic insect control
traits: practical approaches in local environments. Pest Manag. Sci. 66, 100106.
MacIntosh, S.C., McPherson, S.L., Perlak, F.J., Marrone, P.G., Fuchs, R.L., 1990. Purifi-
cation and characterization of Bacillus thuringiensis var. tenebrionis insecticidal proteins pro-
duced in E. coli. Biochem. Biophys. Res. Commun. 170, 665672.
Macrae, T.C., Baur, M.E., Boethel, D.J., Fitzpatrick, B.J., Gao, A.G., Gamundi, J.C.,
Harrison, L.A., Kabuye, V.T., Mcpherson, R.M., Miklos, J.A., Paradise, M.S.,
Toedebusch, A.S., Viegas, A., 2005. Laboratory and field evaluations of transgenic soy-
bean exhibiting high-dose expression of a synthetic Bacillus thuringiensis cry1A gene for
control of Lepidoptera. J. Econ. Entomol. 98, 577587.
242 Kenneth E. Narva et al.
Marra, M.C., Piggott, N.E., Goodwin, B.K., 2012. The impact of corn rootworm protected
biotechnology traits in the United States. AgBioForum 15, 217230.
Masson, L., Schwab, G., Mazza, A., Brousseau, R., Potvin, L., Schwartz, J.L., 2004. A novel
Bacillus thuringiensis (PS149B1) containing a Cry34Ab1/Cry35Ab1 binary toxin specific
for the western corn rootworm Diabrotica virgifera virgifera LeConte forms ion channels in
lipid membranes. Biochemistry 43, 1234912357.
Mattes, O., 1927. Parasitare krankheiten der mehimottenlarven und versuche uber ihre ver-
wendbarkeit als biologisches bekamfungsmittel. Ges Naturw Marburg Schrift. 381.
McDougall, P., 2011. The Cost and Time Involved in the Discovery, Development and
Authorisation of a New Plant Biotechnology Derived Trait. A Consultancy Study for
CropLife International. http://www.croplife.org/PhillipsMcDougallStudy.
McPherson, R.M., MacRae, T.C., 2009. Evaluation of transgenic soybean exhibiting high
expression of a synthetic Bacillus thuringiensis cry1A transgene for suppressing lepidop-
teran population densities and crop injury. J. Econ. Entomol. 102, 16401648.
Mendelsohn, M., Kough, J., Vaituzis, Z., Matthews, K., 2003. Are Bt crops safe? Nat. Bio-
technol. 21, 10031009.
Morin, S., Biggs, R.W., Sisterson, M.S., Shriver, L., Ellers-Kirk, C., Higginson, D.,
Holley, D., Gahan, L.J., Heckel, D.G., Carriere, Y., Dennehy, T.J., Brown, J.K.,
Tabashnik, B.E., 2003. Three cadherin alleles associated with resistance to Bacillus thur-
ingiensis in pink bollworm. Proc. Natl. Acad. Sci. U. S. A. 100, 50045009.
Morse, R.J., Yamamoto, T., Stroud, R.M., 2001. Structure of Cry2Aa suggests an unex-
pected receptor binding epitope. Structure 9, 409417.
Mumm, R.H., 2013. A look at product development with genetically modified crops: exam-
ples from maize. J. Agric. Food Chem. 61, 82548259.
Naimov, S., Weemen-Hendriks, M., Dukiandjiev, S., de Maagd, R.A., 2001. Bacillus thur-
ingiensis delta-endotoxin Cry1 hybrid proteins with increased activity against the Colo-
rado potato beetle. Appl. Environ. Microbiol. 67, 53285330.
Naranjo, S.E., 2010. Impacts of Bt transgenic cotton on integrated pest management.
J. Agric. Food Chem. 59, 58425851.
Narva, K., Siegfried, B., Storer, N., 2013. Transgenic approaches to western corn rootworm
control. In: Vilcinskas, A. (Ed.), Yellow Biotechnology II. Springer, Berlin, pp. 135162.
Noguera, P.A., Ibarra, J.E., 2010. Detection of new cry genes of Bacillus thuringiensis by use of
a novel PCR primer system. Appl. Environ. Microbiol. 76, 61506155.
Ochoa-Campuzano, C., Real, M.D., Martinez-Ramirez, A.C., Bravo, A., Rausell, C.,
2007. An ADAM metalloprotease is a Cry3Aa Bacillus thuringiensis toxin receptor. Bio-
chem. Biophys. Res. Commun. 362, 437442.
Ohba, M., Mizuki, E., Uemori, A., 2009. Parasporin, a new anticancer protein group from
Bacillus thuringiensis. Anticancer Res. 29, 427433.
Oliveira, G.R., Silva, M.C., Lucena, W.A., Nakasu, E.Y., Firmino, A.A., Beneventi, M.A.,
Souza, D.S., Gomes Jr., J.E., de Souza Jr., J.D., Rigden, D.J., Ramos, H.B.,
Soccol, C.R., Grossi-de-Sa, M.F., 2011. Improving Cry8Ka toxin activity towards
the cotton boll weevil (Anthonomus grandis). BMC Biotechnol. 11, 85.
Onstad, D.W., Mitchell, P.D., Hurley, T.M., Lundgren, J.G., Porter, R.P., Krupke, C.H.,
Spencer, J.L., DiFonzo, C.D., Baute, T.S., Hellmich, R.L., Buschman, L.L.,
Hutchison, W.D., Tooker, J.F., 2011. Seeds of change: corn seed mixtures for resistance
management and integrated pest management. J. Econ. Entomol. 104, 343352.
Oppert, B., Kramer, K.J., Beeman, R.W., Johnson, D., McGaughey, W.H., 1997.
Proteinase-mediated insect resistance to Bacillus thuringiensis toxins. J. Biol. Chem.
272, 2347323476.
Pardo-Lopez, L., Munoz-Garay, C., Porta, H., Rodriguez-Almazan, C., Soberon, M.,
Bravo, A., 2009. Strategies to improve the insecticidal activity of Cry toxins from Bacillus
thuringiensis. Peptides 30, 589595.
Bt Crops 243
Pardo-Lopez, L., Soberon, M., Bravo, A., 2013. Bacillus thuringiensis insecticidal three-
domain Cry toxins: mode of action, insect resistance and consequences for crop protec-
tion. FEMS Microbiol. Rev. 37, 322.
Payne, J., Sick, A.J., 1993. Bacillus thuringiensis isolate active against lepidopteran pests, and
genes encoding novel lepidopteran-active toxins. U.S. patent number 5,188,960.
Payne, J.M., Narva, K.E., Uyeda, K.A., Stalder, C.J., Michaels, T.M., 1995. Bacillus thur-
ingiensis isolate PS201T6 toxin. U.S. patent number 5,436,002.
Perez, C., Fernandez, L.E., Sun, J., Folch, J.L., Gill, S.S., Soberon, 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, 1830318308.
Perez, C., Munoz-Garay, C., Portugal, L.C., Sanchez, J., Gill, S.S., Soberon, M., Bravo, A.,
2007. Bacillus thuringiensis ssp. israelensis Cyt1Aa enhances activity of Cry11Aa toxin by
facilitating the formation of a pre-pore oligomeric structure. Cell. Microbiol.
9, 29312937.
Perlak, F.J., Fuchs, R.L., Dean, D.A., McPherson, S.L., Fischhoff, D.A., 1991. Modification
of the coding sequence enhances plant expression of insect control protein genes. Proc.
Natl. Acad. Sci. U. S. A. 88, 33243328.
Petolino, J., Arnold, N., 2009. Whiskers-mediated maize transformation. In: Scott, M.P.
(Ed.), Transgenic Maize. Humana Press, Clifton, NJ, pp. 5967.
Petzold-Maxwell, J.L., Meinke, L.J., Gray, M.E., Estes, R.E., Gassmann, A.J., 2013. Effect of
Bt maize and soil insecticides on yield, injury, and rootworm survival: implications for
resistance management. J. Econ. Entomol. 106, 19411951.
Pigott, C.R., Ellar, D.J., 2007. Role of receptors in Bacillus thuringiensis crystal toxin activity.
Microbiol. Mol. Biol. Rev. 71, 255281.
Porcar, M., Juarez-Perez, V., 2003. PCR-based identification of Bacillus thuringiensis pesti-
cidal crystal genes. FEMS Microbiol. Rev. 26, 419432.
Privalle, L.S., Chen, J., Clapper, G., Hunst, P., Spiegelhalter, F., Zhong, C.X., 2012. Devel-
opment of an agricultural biotechnology crop product: testing from discovery to com-
mercialization. J. Agric. Food Chem. 60, 1017910187.
Qiao, F., Huang, J., Rozelle, S., Wilen, J., 2010. Natural refuge crops, buildup of
resistance, and zero-refuge strategy for Bt cotton in China. Sci. China Life Sci.
53, 12271238.
Radauer, C., Bublin, M., Wagner, S., Mari, A., Breiteneder, H., 2008. Allergens are distrib-
uted into few protein families and possess a restricted number of biochemical functions.
J. Allergy Clin. Immunol. 121, 847852, e847.
Rahman, M.M., Roberts, H.L., Schmidt, O., 2004. The development of the endoparasitoid
Venturia canescens in Bt-tolerant, immune induced larvae of the flour moth Ephestia
kuehniella. J. Invertebr. Pathol. 87, 129131.
Rajagopal, R., Sivakumar, S., Agrawal, N., Malhotra, P., Bhatnagar, R.K., 2002. Silencing
of midgut aminopeptidase N of Spodoptera litura by double-stranded RNA establishes its
role as Bacillus thuringiensis toxin receptor. J. Biol. Chem. 277, 4684946851.
Raybould, A., Higgins, L., Horak, M., Layton, R., Storer, N., De La Fuente, J., Herman, R.,
2012. Assessing the ecological risks from the persistence and spread of feral populations of
insect-resistant transgenic maize. Transgenic Res. 21, 655664.
Reay-Jones, F.P.F., Wiatrak, P., Greene, J.K., 2009. Evaluating the performance of trans-
genic corn producing Bacillus thuringiensis toxins in South Carolina. J. Agric. Urban
Entomol. 26, 7786.
Reed, G.L., Jensen, A.S., Riebe, J., Head, G., Duan, J.J., 2001. Transgenic Bt potato and
conventional insecticides for Colorado potato beetle management: comparative efficacy
and non-target impacts. Entomol. Exp. Appl. 100, 89100.
Ricroch, A.E., 2013. Assessment of GE food safety using -omics techniques and long-term
animal feeding studies. Nat. Biotechnol. 30, 349354.
244 Kenneth E. Narva et al.
Ricroch, A.E., Berge, J.B., Kuntz, M., 2011. Evaluation of genetically engineered crops
using transcriptomic, proteomic and metabolomic profiling techniques. Plant Physiol.
155, 17521761.
Romeis, J., Hellmich, R., Candolfi, M., Carstens, K., De Schrijver, A., Gatehouse, A.,
Herman, R., Huesing, J., McLean, M., Raybould, A., Shelton, A., Waggoner, A.,
2011. Recommendations for the design of laboratory studies on non-target arthropods
for risk assessment of genetically engineered plants. Transgenic Res. 20, 122.
Rosellini, D., 2012. Selectable markers and reporter genes: a well furnished toolbox for plant
science and genetic engineering. Crit. Rev. Plant Sci. 31, 401453.
Roush, R.T., Shelton, A.M., 1997. Assessing the odds: the emergence of resistance to Bt
transgenic plants. Nat. Biotechnol. 15, 816817.
Rule, D.M., Nolting, S.P., Prasifka, P.L., Storer, N.P., Hopkins, B.W., Scherder, E.F.,
Siebert, M.W., Hendrix, W.H., 2014. Efficacy of pyramided Bt proteins Cry1F,
Cry1A.105, and Cry2Ab2 expressed in SmartStax corn hybrids against lepidopteran
insect pests in the northern United States. J. Econ. Entomol. 107, 403409.
Rupar, M.J., Donovan, W.P., Tan, Y., Annette, S., 2000. Bacillus thuringiensis CryET29
compositions toxic to coleopteran insects and Ctenocephalides spp. U.S. patent number
6,093,69.
Saiki, R.K., Gelfand, D.H., Stoffel, S., Scharf, S.J., Higuchi, R., Horn, G.T., Mullis, K.B.,
Erlich, H.A., 1988. Primer-directed enzymatic amplification of DNA with a thermosta-
ble DNA polymerase. Science 239, 487491.
Sanahuja, G., Banakar, R., Twyman, R.M., Capell, T., Christou, P., 2011. Bacillus thur-
ingiensis: a century of research, development and commercial applications. Plant Bio-
technol. J. 9, 283300.
Schnepf, H.E., Whiteley, H.R., 1981. Cloning and expression of the Bacillus thuringiensis
crystal protein gene in Escherichia coli. Proc. Natl. Acad. Sci. U. S. A. 78, 28932897.
Schnepf, E., Crickmore, N., Van Rie, J., Lereclus, D., Baum, J., Feitelson, J., Zeigler, D.R.,
Dean, D.H., 1998. Bacillus thuringiensis and its pesticidal crystal proteins. Microbiol. Mol.
Biol. Rev. 62, 775806.
Schnepf, H.E., Lee, S., Dojillo, J., Burmeister, P., Fencil, K., Morera, L., Nygaard, L.,
Narva, K.E., Wolt, J.D., 2005. Characterization of Cry34/Cry35 binary insecticidal pro-
teins from diverse Bacillus thuringiensis strain collections. Appl. Environ. Microbiol.
71, 17651774.
Sears, M.K., Hellmich, R.L., Stanley-Horn, D.E., Oberhauser, K.S., Pleasants, J.M.,
Mattila, H.R., Siegfried, B.D., Dively, G.P., 2001. Impact of Bt corn pollen on
monarch butterfly populations: a risk assessment. Proc. Natl. Acad. Sci. U. S. A.
98, 1193711942.
Sekar, V., Thompson, D.V., Maroney, M.J., Bookland, R.G., Adang, M.J., 1987. Molecular
cloning and characterization of the insecticidal crystal protein gene of Bacillus thuringiensis
var. tenebrionis. Proc. Natl. Acad. Sci. U. S. A. 84, 70367040.
Sena, J.A., Hernandez-Rodriguez, C.S., Ferre, J., 2009. Interaction of Bacillus thuringiensis
Cry1 and Vip3A proteins with Spodoptera frugiperda midgut binding sites. Appl. Environ.
Microbiol. 75, 22362237.
Shan, G., 2011. Immunoassays in Agricultural Biotechnology. John Wiley & Sons,
Hoboken, NJ.
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. Entomol.
47, 845881.
Shelton, A.M., Olmstead, D.L., Burkness, E.C., Hutchison, W.D., Dively, G., Welty, C.,
Sparks, A.N., 2013. Multi-state trials of Bt sweet corn varieties for control of the corn
earworm (Lepidoptera: Noctuidae). J. Econ. Entomol. 106, 21512159.
Bt Crops 245
Siebert, M.W., Nolting, S.P., Hendrix, W., Dhavala, S., Craig, C., Leonard, B.R.,
Stewart, S.D., All, J., Musser, F.R., Buntin, G.D., Samuel, L., 2012. Evaluation of corn
hybrids expressing Cry1F, Cry1A.105, Cry2Ab2, Cry34Ab1/Cry35Ab1, and Cry3Bb1
against southern United States insect pests. J. Econ. Entomol. 105, 18251834.
Siegel, J.P., 2001. The mammalian safety of Bacillus thuringiensis-based insecticides.
J. Invertebr. Pathol. 77, 1321.
Slaney, A.C., Robbins, H.L., English, L., 1992. Mode of action of Bacillus thuringiensis toxin
CryIIIA: an analysis of toxicity in Leptinotarsa decemlineata (Say) and Diabrotica unde-
cimpunctata howardi Barber. Insect Biochem. Mol. Biol. 22, 918.
Smith, R.H., Hood, E.E., 1995. Agrobacterium-tumefaciens transformation of monocoty-
ledons. Crop Sci. 35, 301309.
Soberon, M., Pardo-Lopez, L., Lopez, I., Gomez, I., Tabashnik, B.E., Bravo, A., 2007. Engi-
neering modified Bt toxins to counter insect resistance. Science 318, 16401642.
Soberon, M., Gill, S.S., Bravo, A., 2009. Signaling versus punching hole: how do Bacillus
thuringiensis toxins kill insect midgut cells? Cell. Mol. Life Sci. 66, 13371349.
Soberon, M., Pardo, L., Munoz-Garay, C., Sanchez, J., Gomez, I., Porta, H., Bravo, A.,
2010. Pore formation by Cry toxins. Adv. Exp. Med. Biol. 677, 127142.
Soberon, M., Lopez-Diaz, J.A., Bravo, A., 2013. Cyt toxins produced by Bacillus thuringiensis:
a protein fold conserved in several pathogenic microorganisms. Peptides 41, 8793.
Stemmer, W.P., 1994. DNA shuffling by random fragmentation and reassembly: in vitro
recombination for molecular evolution. Proc. Natl. Acad. Sci. U. S. A.
91, 1074710751.
Storer, N.P., Kubiszak, M.E., Ed King, J., Thompson, G.D., Santos, A.C., 2012a. Status of
resistance to Bt maize in Spodoptera frugiperda: lessons from Puerto Rico. J. Invertebr.
Pathol. 110, 294300.
Storer, N.P., Thompson, G.D., Head, G.P., 2012b. Application of pyramided traits against
Lepidoptera in insect resistance management for Bt crops. GM Crops Food Biotechnol.
Agric. Food Chain 3, 154162.
Storer, N.P., Thompson, G.D., Head, G.P., 2012c. Application of pyramided traits
against Lepidoptera in insect resistance management for Bt crops. GM Crops Food
3, 154162.
Tabashnik, B.E., Gould, F., 2012. Delaying corn rootworm resistance to Bt corn. J. Econ.
Entomol. 105, 767776.
Tabashnik, B.E., Cushing, N.L., Finson, N., Johnson, M.W., 1990. Field development of
resistance to Bacillus thuringiensis in diamondback moth (Lepidoptera: Plutellidae).
J. Econ. Entomol. 83, 16711676.
Tabashnik, B.E., Liu, Y., Malvar, T., Heckel, D.G., Masson, L., Ferre, J., 1998. Insect resis-
tance to Bacillus thuringiensis: uniform or diverse? Philos. Trans. R. Soc. Lond. B Biol. Sci.
353, 17511756.
Tabashnik, B.E., Huang, F., Ghimire, M.N., Leonard, B.R., Siegfried, B.D.,
Rangasamy, M., Yang, Y., Wu, Y., Gahan, L.J., Heckel, D.G., Bravo, A.,
Soberon, M., 2011. Efficacy of genetically modified Bt toxins against insects with differ-
ent genetic mechanisms of resistance. Nat. Biotechnol. 29, 11281131.
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., Carriere, Y., 2012. Sustained suscep-
tibility of pink bollworm to Bt cotton in the United States. GM Crops Food 3, 194200.
Tabashnik, B.E., Brevault, T., Carriere, Y., 2013. Insect resistance to Bt crops: lessons from
the first billion acres. Nat. Biotechnol. 31, 510521.
Thomas, W.E., Ellar, D.J., 1983. Bacillus thuringiensis var israelensis crystal delta-endotoxin:
effects on insect and mammalian cells in vitro and in vivo. J. Cell Sci. 60, 181197.
246 Kenneth E. Narva et al.
Tomasino, S.F., Leister, R.T., Dimock, M.B., Beach, R.M., Kelly, J.L., 1995. Field perfor-
mance of Clavibacter xyli subsp. cynodontis expressing the insecticidal protein gene cryIA(c)
of Bacillus thuringiensis against European corn borer in field corn. Biol. Control 5, 442448.
U.S. Environmental Protection Agency, 2001a. Biopesticides Registration Action Docu-
mentBacillus thuringiensis (Bt) Plant-Incorporated Protectants. Office of Pesticide
Programs, Biopesticides and Pollution Prevention Division, Washington D.C.
U.S. Environmental Protection Agency, 2001b. A set of scientific issues being considered by
the environmental protection agency regarding: Bt plant-pesticides risk and benefit
assessments: insect resistance management: FIFRA Scientific Advisory Panel Report
No. 2000-07a.
U.S. Environmental Protection Agency, 2007. White Paper on Tier-Based Testing for the
Effects of Proteinaceous Insecticidal Plant-Incorporated Protectants on Non-Target
Arthropods for Regulatory Risk Assessments. Washington D.C. http://www.epa.
gov/opp00001/biopesticides/pips/non-target-arthropods.pdf.
U.S. Environmental Protection Agency, 2011. Updated EPA Review of ABSTCs 2010
Corn Insect Resistance Management Compliance Assurance Program. Memo from J.
Martinez to M. Mendelsohn, Office of Chemical Safety and Pollution Prevention,
Biopesticides and Pollution Prevention Division, Washington D.C.
Vachon, V., Laprade, R., Schwartz, J.L., 2012. Current models of the mode of action of Bacil-
lus thuringiensis insecticidal crystal proteins: a critical review. J. Invertebr. Pathol.
111, 112.
Vaeck, M., Reynaerts, A., Hofte, H., Jansens, S., De Beuckeleer, M., Dean, C., Zabeau, M.,
Montagu, M.V., Leemans, J., 1987. Transgenic plants protected from insect attack.
Nature 328, 3337.
van Frankenhuyzen, K., 2009. Insecticidal activity of Bacillus thuringiensis crystal proteins.
J. Invertebr. Pathol. 101, 116.
van Frankenhuyzen, K., 2013. Cross-order and cross-phylum activity of Bacillus thuringiensis
pesticidal proteins. J. Invertebr. Pathol. 114, 7685.
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 Bacillus thur-
ingiensis protein expressed in transgenic maize. Crop Sci. 45, 931938.
Vidal-Quist, J.C., Castanera, P., Gonzalez-Cabrera, J., 2009. Diversity of Bacillus thuringiensis
strains isolated from citrus orchards in spain and evaluation of their insecticidal activity
against Ceratitis capitata. J. Microbiol. Biotechnol. 19, 749759.
Von Tersch, M.A., Slatin, S.L., Kulesza, C.A., English, L.H., 1994. Membrane-
permeabilizing activities of Bacillus thuringiensis coleopteran-active toxin CryIIIB2 and
CryIIIB2 domain I peptide. Appl. Environ. Microbiol. 60, 37113717.
Wabiko, H., Raymond, K.C., Bulla Jr., L.A., 1986. Bacillus thuringiensis entomocidal
protoxin gene sequence and gene product analysis. DNA 5, 305314.
Walker, K.A., Hellmich, R.L., Lewis, L.C., 2000. Late-Instar European Corn Borer
(Lepidoptera: Crambidae) tunneling and survival in transgenic corn hybrids. J. Econ.
Entomol. 93, 12761285.
Walters, F.S., Stacy, C.M., Lee, M.K., Palekar, N., Chen, J.S., 2008. An engineered chymo-
trypsin/cathepsin G site in domain I renders Bacillus thuringiensis Cry3A active against
Western corn rootworm larvae. Appl. Environ. Microbiol. 74, 367374.
Walters, F.S., deFontes, C.M., Hart, H., Warren, G.W., Chen, J.S., 2010. Lepidopteran-
active variable-region sequence imparts coleopteran activity in eCry3.1Ab, an
engineered Bacillus thuringiensis hybrid insecticidal protein. Appl. Environ. Microbiol.
76, 30823088.
Bt Crops 247
Warren, G.W., 1997. Vegetative insecticidal proteins: novel proteins for control of corn
pests. In: Carozzi, N.B., Koziel, M. (Eds.), Advances in Insect Control, the Role of
Transgenic Plants. Taylors & Francis Ltd., London, pp. 109121.
Wasano, N., Ohba, M., 1998. Assignment of delta-endotoxin genes of the four lepidoptera-
specific Bacillus thuringiensis strains that produce spherical parasporal inclusions. Curr.
Microbiol. 37, 408411.
Widner, W.R., Whiteley, H.R., 1989. Two highly related insecticidal crystal proteins of
Bacillus thuringiensis subsp. kurstaki possess different host range specificities. J. Bacteriol.
171, 965974.
Wu, F., 2006. Mycotoxin reduction in Bt corn: potential economic, health, and regulatory
impacts. Transgenic Res. 15, 277289.
Wu, D., Johnson, J.J., Federici, B.A., 1994. Synergism of mosquitocidal toxicity between
CytA and CryIVD proteins using inclusions produced from cloned genes of Bacillus thur-
ingiensis. Mol. Microbiol. 13, 965972.
Wu, S.J., Koller, C.N., Miller, D.L., Bauer, L.S., Dean, D.H., 2000. Enhanced toxicity of
Bacillus thuringiensis Cry3A delta-endotoxin in coleopterans by mutagenesis in a receptor
binding loop. FEBS Lett. 473, 227232.
Xu, X., Yu, L., Wu, Y., 2005. Disruption of a cadherin gene associated with resistance to
Cry1Ac {delta}-endotoxin of Bacillus thuringiensis in Helicoverpa armigera. Appl. Environ.
Microbiol. 71, 948954.
Yang, Y., Zhu, Y.C., Ottea, J., Husseneder, C., Leonard, B.R., Abel, C., Luttrell, R.,
Huang, F., 2011. Down regulation of a gene for cadherin, but not alkaline phosphatase,
associated with Cry1Ab resistance in the sugarcane borer Diatraea saccharalis. PLoS One
6, e25783.
Ye, W., Zhu, L., Liu, Y., Crickmore, N., Peng, D., Ruan, L., Sun, M., 2012. Mining new
crystal protein genes from Bacillus thuringiensis on the basis of mixed plasmid-enriched
genome sequencing and a computational pipeline. Appl. Environ. Microbiol.
78, 47954801.
Zhang, J., Hodgman, T.C., Krieger, L., Schnetter, W., Schairer, H.U., 1997. Cloning and
analysis of the first cry gene from Bacillus popilliae. J. Bacteriol. 179, 43364341.
Zhang, X., Candas, M., Griko, N.B., Rose-Young, L., Bulla Jr., L.A., 2005. Cytotoxicity of
Bacillus thuringiensis Cry1Ab toxin depends on specific binding of the toxin to the
cadherin receptor BT-R1 expressed in insect cells. Cell Death Differ. 12, 14071416.
Zhang, X., Candas, M., Griko, N.B., Taussig, R., Bulla Jr., L.A., 2006. A mechanism of cell
death involving an adenylyl cyclase/PKA signaling pathway is induced by the Cry1Ab
toxin of Bacillus thuringiensis. Proc. Natl. Acad. Sci. U. S. A. 103, 98979902.
Zhang, S., Cheng, H., Gao, Y., Wang, G., Liang, G., Wu, K., 2009. Mutation of an ami-
nopeptidase N gene is associated with Helicoverpa armigera resistance to Bacillus thur-
ingiensis Cry1Ac toxin. Insect Biochem. Mol. Biol. 39, 421429.
CHAPTER FIVE
Contents
1. Introduction 250
2. Environmental RNAi 252
3. Insect Sensitivity to Environmental RNAi 253
3.1 Coleoptera 253
3.2 Diptera 262
3.3 Lepidoptera 264
3.4 Hemiptera 267
3.5 Other agricultural pests 269
4. Barriers to Delivery and Efficacy in Recalcitrant Species 270
5. Commercial Development of RNAi Actives 276
5.1 Next-generation rootworm-resistant corn 276
5.2 Topical application 278
6. Safety Considerations 280
7. Insect Resistance Management 282
8. Concluding Remarks 284
Acknowledgements 285
References 285
Abstract
Gene suppression via RNA interference (RNAi) provides an alternative strategy for insect
pest management. The ingestion by insects of double-stranded RNAs targeting essen-
tial insect genes can trigger RNAi and lead to growth inhibition, developmental aber-
rations, reduced fecundity, and mortality. This RNAi response is particularly acute in
certain coleopteran species, most notably the western corn rootworm, a devastating
pest impacting corn production in the United States. The development of next-
generation rootworm-protected corn hybrids includes an RNAi-based trait that provides
a mode of action distinct from those of Bacillus thuringiensis insecticidal protein-based
traits currently used for rootworm pest management. Unfortunately, many insect spe-
cies including important lepidoptera and hemiptera pests appear largely recalcitrant in
their response to environmental RNA, suggesting biological barriers that thus far limit
the utility of RNAi for agricultural pest management. This review will highlight recent
efforts to understand the barriers to RNA delivery in recalcitrant insect species, describe
recent advances in the commercial development of insect-protected crops and biolog-
ical insecticides utilizing RNAi, and discuss this strategy in the context of an integrated
pest management approach.
1. INTRODUCTION
Eukaryotic cells possess a conserved pathway by which exogenously
applied and endogenously expressed double-stranded (ds) RNAs direct
the degradation of complementary endogenous messenger RNA (mRNA)
transcripts within a cell resulting in sequence-specific gene suppression. This
phenomenon is referred to as RNA interference (RNAi) (Fire et al., 1998;
Hannon, 2002). In plants and animals, RNAi provides one line of defence
against RNA viruses and foreign dsRNA molecules. Small endogenous
RNAs known as micro RNAs are also processed by a related pathway to
regulate tissue-specific patterns of gene expression primarily via translational
regulation (Bartel, 2009). Long non-coding RNAs also play a prominent
role in the epigenetic regulation of gene expression (Lee, 2012). It is now
clear that far more of the genome is transcribed than previously thought
(Djebali et al., 2012) and that RNA, in addition to being the obligate mes-
senger and facilitator of protein synthesis in the cell, is also a central player in
the regulation of eukaryotic gene expression.
The general mechanism of dsRNA-mediated degradation of mRNA
transcripts is understood (Tomari and Zamore, 2005). Long dsRNAs are
a substrate for RNAse III-like proteins referred to as Dicer or Dicer-like
proteins. Dicer appears to preferentially initiate dsRNA cleavage at the ends
of the dsRNA, making successive cleavages to generate 21- to 24-bp silenc-
ing (si) RNA duplexes (Elbashir et al., 2001). The resulting siRNA duplexes
are loaded into a multiprotein complex called the RNA-induced silencing
complex (RISC) where the passenger (sense) strand is removed and the
guide (antisense) strand remains to target mRNA for silencing. The guide
strand in the RISC enables WatsonCrick base pairing of the complex to
complementary mRNA transcripts and enzymatic cleavage of the target
mRNA by a class of proteins referred to as Argonaute proteins, thereby
preventing mRNA translation. Accordingly, this mechanism of gene sup-
pression is highly sequence specific. The ability to selectively down-regulate
genes via RNAi has proved to be valuable, particularly in insects for which
genetic tools are not readily available to study gene function (Belles, 2010).
RNAi-Mediated Insect Pest Management 251
The surprising observation that ingested dsRNAs can trigger gene suppres-
sion in the nematode, Caenorhabditis elegans, (Timmons and Fire, 1998;
Timmons et al., 2001) offered hope that the oral delivery of dsRNA could
modulate gene expression in other invertebrates, including insects, for the
purpose of pest management.
The ability of herbivorous insects to adapt to insecticide use in agricul-
tural systems presents an ongoing challenge for pest management. Over
20 years ago, more than 500 species of arthropods were documented with
resistance to one or more pesticides (Georghiou et al., 1991); an updated
dataset can be found at http://www.pesticideresistance.com/index.php.
The development of insecticides with new modes of action (MOAs) is a pri-
ority, but so is the implementation of resistance management strategies to
prolong the use of existing insecticides for use in agriculture and public
health (http://www.irac-online.org/). Insect-protected crops, expressing
insecticidal proteins derived from the bacterium, Bacillus thuringiensis (Bt),
represent a significant fraction of the >170 M ha of transgenic crops culti-
vated worldwide ( James, 2013) and provide excellent control of many eco-
nomically important insect pest species. Consequentially, such crops also
impose strong selective pressure on insects to adapt. As is the case for
synthetic- and biological insecticides, alternative MOAs for insect-protected
crops are needed, either because some insect species are not sensitive to Bt
insecticidal proteins or because some have evolved field resistance to effica-
cious Bt proteins (Storer et al., 2010). To that end, RNAi-mediated insect
control represents a significant opportunity.
In 2007, papers from Baum et al. and Mao et al. demonstrated that trans-
genic plants expressing insect-derived dsRNAs could impact the growth and
development of insect herbivores. Corn plants expressing a dsRNA hairpin
that targets the vacuolar ATPase A subunit gene in western corn rootworm
(WCR), Diabrotica virgifera virgifera, caused severe rootworm stunting and
exhibited significant protection from rootworm feeding damage, consistent
with artificial diet feeding assays demonstrating the insecticidal activity of
such dsRNA species (Baum et al., 2007). Arabidopsis plants expressing a
dsRNA hairpin that targets a cytochrome P450 monooxygenase gene in
the cotton bollworm, Helicoverpa armigera, led to decreased bollworm
tolerance to the cotton sesquiterpene aldehyde, gossypol (Mao et al.,
2007), consistent with the proposed role of this enzyme in gossypol
detoxification. The former example has advanced towards commercial
development as a component of next-generation rootworm-protected corn
hybrids (Kupferschmidt, 2013) and will be discussed further in this review.
252 James A. Baum and James K. Roberts
2. ENVIRONMENTAL RNAi
Definitions have been proposed to discuss the various aspects of RNAi
in plants and animals (Huvenne and Smagghe, 2010; Whangbo and Hunter,
2008). Cell-autonomous RNAi refers to the RNAi response that individual
cells carry out when encountering dsRNA, a response that is executed by a
broadly conserved or core RNAi machinery found in eukaryotic cells. Non-
cell autonomous RNAi includes the phenomenon of systemic RNAithe
movement of a silencing signal, presumably siRNA and/or dsRNA, from
cell to cell and from one part of an organism to another. Non-cell autono-
mous RNAi also includes the phenomenon of environmental RNAi which,
as its name suggests, refers to the triggering of RNAi by environmental
exposure to dsRNA by soaking or feeding. Environmental RNAi may or
may not be followed by systemic movement of the silencing signal, perhaps
a key step in determining the biological activity of a dsRNA.
Components of the core RNAi machinery are readily identified
in insect species whose genomes have been sequenced (e.g. Honeybee
Genome Sequencing Consortium, 2006; International Silkworm Genome
Consortium, 2008; The International Aphid Genomics Consortium,
2010; Tomoyasu et al., 2008; Tribolium Genome Sequence Consortium,
2008) and evidence for functional RNAi reported in a wide range of insect
species encompassing the orders Coleoptera, Diptera, Dictyoptera,
Hemiptera, Hymenoptera, Isoptera, Lepidoptera, Neuroptera, and Orthop-
tera. In order to be useful as an insect control agent in agriculture, environ-
mental RNAi must first be operational: dsRNA must be delivered to the
insect either by ingestion or by penetration of the insect cuticle in order
RNAi-Mediated Insect Pest Management 253
Apis mellifera Vitellogenin Second Natural diet Nonspecific (5003000 ppm) Yes Nunes and Simoes
instars (2009)
Solenopsis invicta PBAN/pyrokinin Fourth Artificial diet Yes (1000 ppm) ND Vander Meer and
instars Choi (2013)
GNBP Workers Artificial diet Yes (200 ppm) ND Zhao and Chen
(2013)
Isoptera
Reticulitermes Cellulase Workers Paper discs Yes (5.1 g/cm2) Yes Zhou et al. (2008)
flavipes
Hexamerin Workers Paper discs Yes (2.2 g/cm )2
Yes Zhou et al. (2008)
Lepidoptera
Chilo infuscatellus CiHR3 moulting Third Corn kernels Yes (250 ppm) Yes Zhang et al.
factor instars (2012c)
Epiphyas Carboxylesterase Third Droplet No (4000 ppm) Yes Turner et al.
postvittana instars (2006)
Pheromone bp Third Droplet No (4000 ppm) Yes Turner et al.
instars (2006)
Continued
Table 5.1 Sensitivity of insect species to ingested dsRNAscont'd
LC50 or
Target gene Mortality or (concentration mRNA
Organism product Stage Assay stunting tested) silencing Reference
Helicoverpa armigera CYP6AE14 Third Transgenic Yes ND Yes Mao et al. (2007)
instars plant
GST Third Transgenic No ND Yes Mao et al. (2007)
instars plant
CYP6AE14 Third Transgenic Yes ND Yes Mao et al. (2011,
instars plant 2013)
AchE receptor Neonates Artificial diet Yes (0.35 ppm) Yes Kumar et al. (2009)
AchE receptor Neonates Leaf tissue Yes (0.35 ppm) Kumar et al. (2009)
Ecdysone Second Transgenic Yes ND Yes Zhu et al. (2012)
receptor EcR instars plant
Ecdysone Third Artificial diet Yes ND Yes Zhu et al. (2012)
receptor EcR instars (Ec)b
HaHR3 Third Transgenic Yes ND Yes Xiong et al. (2013)
moulting factor instars plant
HaHR3 Third Artificial diet Yes ND Yes Xiong et al. (2013)
moulting factor instars (Ec)
CYP6B6 Third Artificial diet Yes ND Yes Zhang et al.
instars (Ec) (2013c)
Ultraspiracle Third Artificial diet Yes (1000 ppm) Yes Yang and Han
protein, EcR instars (2014)
Manduca sexta V-ATPase E Neonates Artificial diet Yes 11 ppm Yes Whyard et al.
(2009)
Ostrinia nubilalis Chitinase Neonates Artificial diet Yes (2500 ppm) Yes Khajuria et al.
(2010)
Plutella xylostella CYP6BG1 Fourth Droplet (800 ppm) Yes Bautista et al.
instars (2009)
Rieske protein Second Leaf tissue Yes (6 g/cm2) Yes Gong et al. (2011)
instars
AchE receptor Second Leaf tissue Yes 53.7 ppm - Gong et al. (2013)
instars
Sesamia nonagriodes JH esterase JHER First to Artificial diet No ND Yes Kontogiannatos
sixth instars (Ec) et al. (2013)
Spodoptera exigua Chitin synthase A Neonates Artificial diet Yes ND Yes Tian et al. (2009)
(Ec)
1 integrin Fourth Leaf tissue Yes 100200 ppm Surakasi et al.
subunit instars (2011)
Ecdysone Second Transgenic Yes ND Yes Zhu et al. (2012)
receptor EcR instars plant
Continued
Table 5.1 Sensitivity of insect species to ingested dsRNAscont'd
LC50 or
Target gene Mortality or (concentration mRNA
Organism product Stage Assay stunting tested) silencing Reference
Spodoptera litura Aminopeptidase Neonates Artificial diet No ND No Rajagopal et al.
N (2002)
Spodoptera Allatostatin C Fifth instars Droplet NA (600 ppm) Yes Griebler et al.
frugiperda (2008)
Allototropin 2 Fifth instars Droplet NA (600 ppm) Yes Griebler et al.
(2008)
SfT6 serine Fourth Droplet NA (600 ppm) Yes Rodrguez-
protease instars Cabrera et al.
(2010)
Orthoptera
Gryllus bimaculatus Sulfakinins Adults Droplet NA (100600 ppm) Meyering-Vos and
Muller (2007)
Locusta migratoria Multiple targets Fourth Artificial diet No (240 ppm) No Luo et al. (2013)
instars
Schistocerca gregaria Tubulin, Adults Artificial diet No ND No Wynant et al.
GAPDH (2012)
Estimated from sample overlay assays in which 20 L samples are infused into 200 L artificial diet.
a
targets). While the suppression of certain gene transcripts did not result in a
phenotypic response (Baum et al., 2007, supplemental figure 3), it is clear
that the number of specific gene targets available for successful environmen-
tal RNAi is large. Sectioning of the V-ATPase A coding region into six
300 bp dsRNAs did not reveal dramatic differences in efficacy, suggesting
that a single dsRNA of this size provides a reasonable sampling of target
knockdown efficacy and phenotype. One of the targets identified by
Baum et al. (2007), a Snf7 ortholog, was selected for a more detailed study
of the RNAi response (Bolognesi et al., 2012). Snf7 dsRNAs shorter than
50 bp exhibited dramatically reduced activity in the corn rootworm feeding
assay. This study employed a single Snf7 27-mer sequence embedded in
neutral sequences of varying length, ruling out the possibility that differences
in siRNA composition accounted for differences in activity. An in situ study
of RNA uptake further demonstrated that a Cy3-labelled 240 bp Snf7
dsRNA was taken up by rootworm midgut epithelial cells while a Cy3-
labelled 21 bp Snf7 siRNA was not, corroborating the size dependency of
the RNAi response observed in feeding assays (Bolognesi et al., 2012). Like-
wise, injection studies with corn rootworm larvae have demonstrated the
inability of siRNAs to produce an RNAi response leading to rootworm
mortality (Khajuria et al., 2013). Finally, the RNAi response in corn
rootworm appears to be systemic as judged by qPCR analysis of Snf7
mRNA transcripts in isolated midgut and cadaver tissues (Bolognesi
et al., 2012).
Elements of the rootworm RNAi response can be found in another cole-
opteran species, the red flour beetle T. castaneum, which has become a model
system for studying systemic RNAi in insects (Miller et al., 2012). Injection
studies with a transgenic gfp-Tribolium line permitted visualization of silenc-
ing as suppression of green fluorescent protein (gfp) fluorescence. The
RNAi response was observed to be dose-dependent, systemic, and likewise
dependent on the size of the dsRNA. Injection of 60 ng of a 520-bp gfp
dsRNA/last instar larva resulted in detectable silencing of gfp, demonstrating
that this coleopteran species is very sensitive to systemic RNAi. The effi-
ciency of systemic RNAi appears to drop off with shorter dsRNA molecules
although the precise breakpoint was not determined. Adopting a strategy
similar to that used for characterization of the corn rootworm Snf7 dsRNA
(Bolognesi et al., 2012), a 30-bp gfp dsRNA fused to a 30-bp neutral Ultra-
bithorax (Ubx) dsRNA was shown to be more effective than the 30-bp gfp
dsRNA alone in triggering a systemic RNAi response, presumably due to
less efficient cellular uptake of the smaller dsRNA. Comparing a series of
262 James A. Baum and James K. Roberts
3.2. Diptera
Belles (2010) provides some historical context for the pioneering RNAi
studies conducted with the fruit fly, Drosophila melanogaster, a model system
for understanding both the mechanism of RNAi and its role in mediating
antiviral immunity in invertebrates (Nayak et al., 2013). Reverse genetic
studies of gene function have been enabled through the use of cultured cells,
employing both genome-wide (Boutros and Ahringer, 2008; Boutros et al.,
2004) and pathway-specific screens (e.g. Clemens et al., 2000). Functional
studies of genes involved in Drosophila development have relied on micro-
injection of dsRNA into embryonic tissue (e.g. Kennerdell and Carthew,
1998; Koizumi et al., 2007; Pilot et al., 2006). Libraries of transgenic Dro-
sophila lines containing conditionally expressed dsRNA hairpins have been
generated for use in whole-animal RNAi screens (e.g. Dietzl et al., 2007).
Abundant tools for both in vivo and in vitro RNAi studies in Drosophila, as
well as relevant literature, can be found at http://www.flyrnai.org/.
Drosophila appears deficient in systemic RNAi when confronted with
endogenously expressed dsRNA hairpins (Roignant et al, 2003), but is
RNAi-Mediated Insect Pest Management 263
3.3. Lepidoptera
Terenius et al. (2011) provided a useful overview of the status of RNAi stud-
ies in lepidopteran species and noted that, with respect to oral delivery, gene
suppression only appears to be successful when high concentrations of
dsRNA are provided in the diet. In the first reported example, Turner
et al. (2006) achieved a significant suppression of several target genes in
the brown apple moth, Epiphyas postvittana (Walker), by droplet feeding a
4000 ppm dsRNA solution. Subsequently, silencing via oral delivery of
dsRNA was reported in a wide range of lepidopteran species including
the tobacco hornworm, Manduca sexta, diamondback moth (DBM), Plutella
xylostella, beet armyworm, Spodoptera exigua, fall armyworm, Spodoptera
frugiperda, European corn borer, Ostrinia nubilalis, sugarcane stem borer,
Chilo infuscatellus, and the cotton bollworm, H. armigera (Bautista et al.,
2009; Gong et al., 2011; Khajuria et al., 2010; Mao et al., 2007;
Rodrguez-Cabrera et al., 2010; Surakasi et al., 2011; Tian et al., 2009;
Whyard et al., 2009; Xiong et al., 2013; Zhang et al., 2012c; Zhu
et al., 2012).
For many early RNAi studies, the object was not to kill an insect but to
selectively down-regulate a gene to study its function in a metabolic or
developmental process (Belles, 2010). For example, environmental RNAi
was used to examine the role of two chitinase genes in regulating chitin con-
tent in the peritrophic matrix of the European corn borer (Khajuria et al.,
2010), to demonstrate that the cytochrome P450 gene, CYPBG1, in
DBM is involved in larval resistance to the insecticide permethrin
(Bautista et al., 2009), to study the role of the beet armyworm 1 subunit
integrin (Se1) in development and cellular immunity (Surakasi et al.,
2011), and to demonstrate that a serine protease gene in the fall armyworm
RNAi-Mediated Insect Pest Management 265
3.4. Hemiptera
RNAi techniques have been used successfully in a wide variety of hemip-
teran species encompassing phloem feeders such as aphids and piercing
sucking insects such as plant bugs (Li et al., 2013b). Since many hemipteran
species are relatively small and fragile as nymphs, oral delivery of dsRNAs for
gene silencing has been an attractive alternative to microinjection. As was
the case with lepidopterans, early studies were not necessarily focused on
killing insects but on studying gene function (e.g. Belles, 2010; Paim
et al., 2013). Across all studies with hemipterans, the dietary concentrations
of dsRNA required for silencing and/or lethal phenotypes vary widely, even
between studies with the same organism, but tend to be at least three orders
of magnitude higher than effective concentrations used with sensitive cole-
opteran species. Examples of gene silencing following ingestion of dsRNA
include the nitrophorin 2 (Np2) gene in the triatomine bug, Rhodnius prolixus
(Araujo et al., 2006), the aquaporin 1 (ApAQP1) gene in the pea aphid,
Acyrthosiphon pisum (Shakesby et al., 2009), the vacuolar ATPase subunit
E gene in A. pisum (Whyard et al., 2009), the trehalose phosphate synthase
(tps) gene in brown planthopper, Nilaparvata lugens (Chen et al., 2010), the
gap gene hunchback in A. pisum (Mao and Zeng, 2012), and the vacuolar
ATPase subunit E gene in N. lugens (Li et al., 2011a). Drawing from a list
of efficacious gene targets identified for the WCR, Upadhyay et al.
(2011) reported silencing of the vacuolar ATPase subunit Aand ribosomal
protein L9 genes in the whitefly, Bemisia tabaci, as well as mortality with
LC50 values in the 311 ppm range. Focusing on gene targets that are highly
or specifically expressed in the midgut, Wuriyanghan et al. (2011) demon-
strated target gene suppression and lethality in the potatotomato psyllid,
Bactericerca cockerelli, when dsRNAs were presented at high concentrations
(5001000 ppm) in a 15% sucrose diet. Both studies also reported lethality
upon ingestion of siRNAs, an effect that has not been reported in
268 James A. Baum and James K. Roberts
gene silencing and 100% mortality after 7 days of feeding on plants (Mutti
et al., 2006), suggesting that the delivery of dsRNA via plant expression was
simply inadequate for control of this species. A similar outcome was reported
using transgenic rice plants expressing dsRNAs targeting three midgut-
expressed genes in N. lugens (Zha et al., 2011). In this study, significant
reductions in target transcript levels (as high as 73% for the Nltry gene) were
detected 24 days after feeding but no lethal phenotype was observed.
Comparisons across related studies reveal marked variation in the hemip-
teran response to environmental RNAi. For example, Chen et al. (2010)
reported significant silencing of the trehalose phosphate synthase (tps) gene
in N. lugens at dsRNA diet concentrations of 500 ppm but no silencing at
concentrations of <100 ppm while Li et al. (2011a) reported significant
silencing of the vacuolar ATPase subunit E gene in N. lugens at dsRNA diet
concentrations of only 50 ppm. Silencing of the same essential gene in two
distantly related hemipteran species can apparently yield different pheno-
types: a dsRNA silencing the vacuolar ATPase subunit E gene in
A. pisum yielded a reported LC50 of 3.4 ppm (Whyard et al., 2009) while
a dsRNA silencing the N. lugens ortholog caused no phenotype (Li et al.,
2011a). Moreover, different groups working with the same insect species
and gene target have reported conflicting outcomes. Whereas Whyard
et al. (2009) reported oral efficacy with a dsRNA targeting the vacuolar
ATPase subunit E gene in A. pisum and Mutti et al. (2006) reported lethality
following injection of C002 dsRNA in A. pisum, Christiaens et al. (2014)
failed to observe a phenotypic response with either dsRNA and delivery
method in this species. In this case, the decision to deploy a long COO2
dsRNA rather than pre-diced siRNAs could have influenced the outcome
of the injection experiment. Other factors that could account for these dis-
parate results include insect colony source and health, viral load, life stage,
dsRNA stability in different diets, and target gene propensity for silencing.
2013a; Wang et al., 2011; Yao et al., 2013; Zhang et al., 2013b). Contrary to
this view, the results of dsRNA screens with sensitive coleopteran species
yielding high hit rates among housekeeping gene targets using low dietary
concentrations of dsRNA support the argument that biological barriers to
RNAi, not target gene selection, are the problem. Likewise, the need to pro-
vide dsRNAs targeting obviously essential genes (e.g. the V-ATPase subunit
genes) at dietary concentrations more than 1000-fold higher than are
required for an RNAi phenotype in sensitive coleopteran species suggests
that there are great inefficiencies in triggering and sustaining an RNAi
response in recalcitrant species. It seems unlikely that the selection of alter-
native gene targets will close this efficacy gap. The absence of RNAi-induced
lethal phenotypes in recalcitrant species may simply reflect the degree to
which gene silencing actually occurs. While many studies report target gene
silencing upon ingestion of dsRNA, the level of transcript knockdown sel-
dom exceeds 60% and silencing is frequently transient (Huvenne and
Smagghe, 2010; Li et al., 2013b). In the case of sensitive coleopteran species,
target gene knockdown is frequently 90% (Baum et al., 2007; Bolognesi
et al., 2012; Rangasamy and Siegfried, 2012; Zhu et al., 2011). The typical
twofold reduction in mRNA transcript levels observed in many species may
fall well within the range of normal variation observed in biological systems
and, in any case, may not manifest immediately during the course of a feeding
assay since the target protein itself needs to turnover first.
If we set aside the question of target gene selection and dsRNA design,
the barriers to efficient RNAi in recalcitrant insects can be pictured as dis-
creet steps in the mechanism of action of RNAi. DsRNAs, whether applied
topically to an insect diet or expressed in transgenic plants, must be able to
survive long enough to be taken up by insect cells. The insect midgut with its
brush border membrane topography (Chapter 1) represents a large surface
area that facilitates the absorption of nutrients and is the presumed site of
entry for dsRNAs in most studies. It is not clear to what extent the
peritrophic matrix in lepidopteran- and coleopteran midguts (Hegedus
et al., 2009; Lehane, 1997) or the perimicrovillar membrane in hemipteran
midguts (Silva et al., 2004) represents a physical barrier to dsRNA delivery.
Upon safe arrival at the gut membrane surface, dsRNAs must be efficiently
taken up by epithelial cells and delivered to the intracellular RNAi machin-
ery. If the route of cellular uptake involves endocytosis, then release or
escape from the endosome becomes a critical step for delivery to the cyto-
plasm (Varkouhi et al., 2011). The core RNAi machinery, being agnostic
with respect to the source of the dsRNA, processes the dsRNA to generate
siRNARISC complexes that can trigger a silencing event, provided
272 James A. Baum and James K. Roberts
transcripts of the gene target are present in that host cell. It is likely that
expression of the RNAi core machinery, including Dicer and Argonaute
proteins, varies among insect cell types and species, contributing to differ-
ences in the efficiency of cell-autonomous RNAi. Factors impacting
RNA amplification and/or systemic movement of the silencing signal from
the site of cellular entry represent additional barriers to RNAi efficiency and
the generation of lethal phenotypes.
The alkaline nature of the lepidopteran midgut (Dow, 1992) could cer-
tainly represent one barrier to RNA delivery. Across insect orders, the pres-
ence of midgut and salivary nucleases that can rapidly degrade dsRNA
molecules likely constitutes a more significant barrier (Arimatsu et al.,
2007a,b; Christiaens et al., 2014; Furusawa et al., 1993; Liu et al., 2013;
Luo et al., 2013; Rodrguez-Cabrera et al., 2010; Terenius et al., 2011;
Wynant et al., 2014). In the case of hemipteran species such as the tarnished
plant bug, L. lineolaris, which engage in extra-oral digestion of plant material
prior to ingestion, the presence of dsRNAses in salivary secretions represents
a significant barrier to dsRNA survival (Allen and Walker, 2012). Likewise,
Christiaens et al. (2014) provided evidence for dsRNA degradation by
dsRNAses in both pea aphid salivary secretions and haemolymph. No less
than four different midgut-expressed dsRNAses were observed in the desert
locust, S. gregaria (Wynant et al., 2014), a species that is recalcitrant to
ingested dsRNA but exhibits a potent systemic RNAi response to injected
dsRNA (Wynant et al., 2012). Knockdown experiments implicated at least
one of these nucleases in dsRNA degradation in the midgut. Likewise, Luo
et al. (2013) reported the presence of potent dsRNAses in midgut fluid from
the migratory locust, L. migratoria, and concluded that rapid degradation of
dsRNA in the midgut explains the recalcitrant nature of this species to die-
tary dsRNA. Given the sensitivity of both locust species to injected
dsRNAs, it would be worth understanding the extent to which dsRNAses
are also expressed in the haemolymph. Finally, the importance of a gut envi-
ronment hostile to dsRNA is supported by our own observation that nucle-
ases found in gut extracts from S. frugiperda, a lepidopteran species, are far
more active in degrading dsRNAs than are nucleases found in corn
rootworm gut extracts (Fig. 5.1), particularly at an alkaline pH more typical
of the lepidopteran midgut.
Inefficient uptake by midgut epithelial cells represents the next barrier to
efficacious RNAi-mediated insect control. The mechanism of dsRNA
uptake by insect gut epithelial cells has not been defined, even for highly
sensitive species such as the WCR. In the nematode, C. elegans, several genes
RNAi-Mediated Insect Pest Management 273
Figure 5.1 dsRNAs are rapidly degraded when treated with midgut extracts from the
lepidopteran species, Spodoptera frugiperda, but remain intact when treated with mid-
gut extracts from the western corn rootworm, Diabrotica virgifera virgifera. A 400-bp
V-ATPase A dsRNA (10 g, see arrow) was incubated at 23 C with 50 g of total protein
extracted from isolated midgut tissue. Incubations were carried out in 100 L volumes in
either 25 mM Tris HCl (pH 7.4) or 25 mM sodium carbonate (pH 10.5). Aliquots (20 L)
were drawn at the times indicated, the digestions were quenched by ethanol precipi-
tation, and the recovered dsRNAs were resolved by agarose gel electrophoresis.
MW Invitrogen molecular weight standards.
required for environmental RNAi and systemic spread have been identified.
The Sid-1 gene is required for systemic movement of the silencing signal
between cells and is also required for environmental RNAi by gut epithelial
cells. The Sid-1 gene encodes a transmembrane protein that acts as a
dsRNA-selective and dsRNA-gated channel to passively traffic dsRNAs
between cells (Feinberg and Hunter, 2003; Jose et al., 2009; Shih and
Hunter, 2011; Shih et al., 2009; Winston et al., 2002). The Sid-2 gene
encodes a membrane protein exclusively localized to intestinal cells that is
required for dsRNA uptake from the intestinal lumen (McEwan et al.,
2012; Winston et al., 2007). Environmental RNAi therefore requires coor-
dination between the SID-1 and SID-2 proteins. A model has been pro-
posed whereby dsRNAs are imported from the intestinal lumen by
SID-2 via endocytosis and released from the internalized vesicles via passive
274 James A. Baum and James K. Roberts
movement through the SID-1 channel (McEwan et al., 2012; Whangbo and
Hunter, 2008). While putative Sid-1-like genes have been identified in some
insects, there are no reports of Sid-2-like genes in insect species whose
genomes have been completely sequenced (e.g. Tomoyasu et al., 2008;
Xu and Han, 2008; Zha et al., 2011).
In Drosophila, an insect lacking Sid-1-like genes, evidence has been pres-
ented for an energy-dependent endocytic mechanism of dsRNA uptake by
cultured (S2) cells, apparently mediated by pattern-recognition scavenger
receptors (Saleh et al., 2006; Ulvila et al., 2006). A genomic screen identified
23 genes required for RNAi in S2 cells including many components of the
endocytic pathway involved in vesicle trafficking and protein transport or
sorting. Subsequently, orthologs of these genes were selected for RNAi
screens in C. elegans, resulting in the identification of 10 genes whose silenc-
ing disrupted the RNAi response triggered by feeding (Saleh et al., 2006). As
noted by Hinas et al. (2012), this screen could not distinguish among knock-
outs that affected dsRNA uptake by intestinal cells, systemic RNAi, or the
autonomous cellular RNAi machinery. In a subsequent study, mutations in
several of the endocytic pathway genes required for cellular dsRNA uptake
in Drosophila rendered flies that were hypersensitive to viral infection and
unable to mount a systemic antiviral response (Saleh et al., 2009). Further
support for the role of endocytic pathways in systemic RNAi proceeded
from the characterization of sid-3 and sid-5 mutants in C. elegans, mutants
that are defective in systemic RNAi but are otherwise RNAi competent.
Sid-3 encodes a conserved tyrosine kinase, homologs of which are known
to associate with endocytic vesicles, and is required for the import of dsRNA
into cells ( Jose et al., 2012). Sid-5 encodes an endosome-associated protein
required for the export of dsRNAs from cells (Hinas et al., 2012). Three
other mutants identified in RNAi screens, rsd-2, rsd-3, and rsd-6, are insen-
sitive to ingested dsRNA targeting germline-expressed genes but remain
sensitive to injected dsRNA (Tijsterman et al., 2004). Other mutants iden-
tified in this screen, rsd-8 and rsd-4, were subsequently shown to be alleles of
sid-1 and sid-2, respectively, the former already identified as required for sys-
temic RNAi. While the details are not yet understood, we expect that sys-
temic RNAi in insects will prove to be equally complex and that Sid-1-like
genes will not prove to be the sole determinant of systemic RNAi. Indeed,
it is not entirely clear that the Sid-1-like genes identified to date in insects
are even involved in systemic RNAi. A phylogenetic analysis of Sid-1-like
proteins in Tribolium suggests that these proteins may not be orthologous to
Sid-1, but rather to the C. elegans Tag-130 protein that is not involved in
RNAi-Mediated Insect Pest Management 275
Figure 5.2 Transgenic plants expressing both the Cry3Bb1 protein and corn rootworm
(CRW)-specific dsRNAs show excellent protection from rootworm feeding damage. Root
damage ratings (03 Iowa scale) are averaged across nine locations from four states (IA,
IL, IN, NE). Standard error bars are shown.
RNAi-Mediated Insect Pest Management 277
40%
% Beetle emergence
30%
20%
10%
0%
Negative Cry3Bb1 V-ATPase V-ATPase
control + Cry3Bb1
Figure 5.3 Transgenic plants expressing the CRW V-ATPase A dsRNA exhibit superior
control of CRW adult beetle emergence in growth chamber assays. Standard error bars
are shown.
278 James A. Baum and James K. Roberts
Figure 5.4 The mechanism and delivery strategies for RNA interference. RNAi therapeu-
tics (e.g. siRNA) can be internalized into cells via different delivery vehicles.
Figure reproduced with permission from Zhou et al. (2013).
RNAi-Mediated Insect Pest Management 279
in this case Drosophila (Whyard et al., 2009). Zhang et al. (2010) demon-
strated that chitosandsRNA complexes were stabilized in gel-based diets
for delivery to mosquito larvae, but the actual impact of chitosan on dsRNA
delivery to or uptake by mosquito gut cells was not evaluated in this study.
More recently, He et al. (2013) reported that fluorescent nanoparticles
(FNPs) comprised of a perylene-3,4,9,10-tetracarboxydiimide chromo-
phore core, a polyphenylene dendrimer inner shell, and a cationic polymer
outer shell can serve as a delivery vehicle for both DNA and dsRNA. FNPs
coated with a dsRNA targeting the chitinase-like CHT10 gene were fed at a
dietary concentration of 250 ppm to the lepidopteran species, O. furnacalis,
and shown to be more effective than dsRNA alone in arresting larval
development.
A formulation of chemically modified AChE2 siRNA complexed with
chitosan was shown to be toxic to DBM larvae in leaf disc assays and in a field
trial (Gong et al., 2013). Larval mortality (%) in the field appeared to peak at
5 days post-treatment with the siRNA formulation (50, 100, and 200 ppm
treatments), perhaps reflecting foliar degradation of the siRNA active or
alternatively an increase in larval infestation leading to a higher recovery
of survivors. The high sensitivity of CPB larvae to ingested dsRNAs pro-
vides an opportunity to explore the development of dsRNA formulations
for use as foliar sprays in insect control. Zhu et al. (2011) evaluated the
use of E. coli-expressed dsRNA to control CPB, demonstrating consistent
target gene knockdown and high mortality of CPB larvae when fed potato
leaves treated with either bacterial-expressed dsRNA or in vitro-
synthesized dsRNA.
The cost of goods weighs heavily on the development of RNAi spray
formulations, including the cost of dsRNA production as well as any req-
uisite delivery agents. The first outdoor trial with an RNAi product was
conducted by Hunter et al. (2010) to assess the efficacy of a dsRNA product
(Remebee-IAPV) in reducing Israeli acute paralytic disease in honey bees,
one of many viral diseases that afflict honey bee colonies worldwide. In this
case, dsRNA was provided to honey bees in a sugar solution with the aim of
blocking virus replication and transmission. Formulations that improve
RNA delivery and insecticidal activity while enhancing environmental sta-
bility could enable a more favourable cost of goods (through lower use rates)
and the development of economically viable topical applications for insect
control. Factors that likely impact the environmental stability of dsRNA are
not so different from those impacting the stability of synthetic insecticides
and include UV damage from exposure to sunlight, microbial degradation
280 James A. Baum and James K. Roberts
6. SAFETY CONSIDERATIONS
Safety assessments for the use of RNAi strategies in insect pest man-
agement include both the evaluation of environmental safety and food and
feed safety. The impact of dsRNAs on non-target organisms, including
beneficial insects, and the persistence of applied dsRNAs in the environ-
ment, are also addressed during the registration of Bt-based bioinsecticides
and insect-protected crops. The current safety assessment approach to
biotechnology-derived plants is appropriate for assessing the safety of plants
produced using RNA-mediated gene regulation (Codex, 2003; EFSA,
2006; FDA, 1994; Parrott et al., 2010; Petrick et al., 2013; Redenbaugh
et al., 1992), and the current environmental risk assessment approach can
be applied to characterize risk for RNA-based technologies (Auer and
Frederick, 2009; ILSI-CERA, 2011).
RNAi technology has the potential for high specificity towards target
pest organisms, even more so than Bt-based technology (Bachman et al.,
2013; Baum et al., 2007; Whyard et al., 2009). Larval feeding assays with
orthologous CPB- and WCR dsRNAs demonstrated that the accumulation
of base pair mismatches between the dsRNA sequence and target gene
sequence results in reduced insecticidal activity (Baum et al., 2007). For
example, a V-ATPase A dsRNA from WCR sharing 83% sequence identity
with the CPB V-ATPase A coding region showed >10-fold lower activity
against CPB larvae than did the orthologous CPB V-ATPase A dsRNA.
Note that this residual activity towards CPB was expected because the
WCR dsRNA sequence still retained stretches of >21 bp sequence identity
with the CPB V-ATPase coding region, a result that is entirely consistent
with the work of Bolognesi et al. (2012) and Bachman et al. (2013). The
case for sequence specificity is further supported by the work of Whyard
et al. (2009), who demonstrated that four Drosophila species could be selec-
tively controlled using dsRNAs that target the divergent 30 untranslated
region of the Tub23C tubulin gene, a region lacking 21 bp matches among
the four species. Other factors, including the barriers to RNAi described in
RNAi-Mediated Insect Pest Management 281
glycans that are also found along the intestinal tract of mammalian herbivores
(Vandenborre et al., 2009). Without complementary traits for resistance
management, the commercialization of durable RNAi-based hemipteran
control traits may prove elusive.
8. CONCLUDING REMARKS
The mechanism of gene suppression manifested as RNAi provides an
MOA unique among insecticidal agents. In addition, the potential for highly
tailored insecticidal specificity afforded in part by the mechanism of RNAi
differentiates it from other insect pest management strategies. Successful
development of RNAi actives, either expressed in plants or applied topically
as biological insecticide sprays, would provide growers with an important
tool for sustainable insect pest management.
Despite the exceptional sensitivity of corn rootworm larvae to ingested
V-ATPase- or Snf7 dsRNAs, transgenic corn plants expressing these
dsRNAs do not provide complete protection from rootworm feeding dam-
age in the field, presumably because of the slow speed-to-kill typical of
dsRNA actives. Transgenic corn plants combining the Snf7 RNAi trait with
the Cry3Bb1 gene, however, do provide superior control of corn rootworm
larvae and emerging adults in the field. Accordingly, stacking the Snf7 RNAi
trait with suitable Bt traits in corn offers the best opportunity for efficacious
and durable control of corn rootworm species. Looking towards the future,
it seems likely that the slow speed-to-kill exhibited by insecticidal dsRNAs
will affect how these agents are deployed commercially.
It is now firmly established that insect species differ greatly in their
response to environmental RNA. Most notably, the lepidopteran and
hemipteran species studied to date are far less sensitive to ingested dsRNAs
than are the responsive coleopteran species exemplified by the corn
rootworm. This differential sensitivity to environmental RNAi helps inform
the environmental safety assessment of RNAi for management of coleop-
teran pests. It will also dictate the approach taken to exploit the phenomenon
of RNAi for insect pest management. In order to achieve parity with the
WCR Snf7 dsRNA active, for example, dsRNAs for lepidopteran- or
hemipteran pests will need at least a 1000-fold improvement in oral activity.
Given the recalcitrant nature of most lepidopteran and hemipteran species to
environmental RNAi and the minimal impact target gene selection will
likely have on closing this efficacy gap, a clear path towards development
of RNAi-based plant traits for lepidopteran- and hemipteran pest
RNAi-Mediated Insect Pest Management 285
ACKNOWLEDGEMENTS
The authors wish to thank Greg Heck, Sergey Ivashuta, Steve Levine, Bill Moar, Jay Petrick,
Gerrit Segers, and Greg Watson for their thorough review of the manuscript.
REFERENCES
Akhtar, S., 2009. Oral delivery of siRNA and antisense oligonucleotides. J. Drug Target.
17, 491495.
Allen, M.L., Walker III, W.B., 2012. Saliva of Lygus lineolaris digests double stranded
ribonucleic acids. J. Insect Physiol. 58 (3), 391396.
Araujo, R.N., Santos, A., Pinto, F.S., Gontijo, N.F., Lehane, M.J., Pereira, M.H., 2006.
RNA interference of the salivary gland nitrophorin 2 in the triatomine bug Rhodnius
prolixus Hemiptera, Reduviidae by dsRNA ingestion or injection. Insect Biochem.
Mol. Biol. 36 (9), 683693.
Arimatsu, Y., Furuno, T., Sugimura, Y., Togoh, M., Ishihara, R., Tokizane, M., Kotani, E.,
Hayashi, Y., Furusawa, T., 2007a. Purification and properties of double-stranded RNA
degrading nuclease, dsRNase, from the digestive juice of the silkworm, Bombyx mori.
J. Insect Biotechnol. Sericol. 76, 5762.
286 James A. Baum and James K. Roberts
Arimatsu, Y., Kotani, E., Sugimura, Y., Furusawa, T., 2007b. Molecular characterization of a
cDNA encoding extracellular dsRNase and its expression in the silkworm, Bombyx mori.
Insect Biochem. Mol. Biol. 37, 176183.
Aronstein, K., Oppert, B., Lorenzen, M., 2011. RNAi in agriculturally-important arthro-
pods. In: Grabowski, P.P. (Ed.), RNA Processing. In Tech, Shanghai, China,
pp. 157180.
Auer, C., Frederick, R., 2009. Crop improvement using small RNAs: applications and pre-
dictive ecological risk assessments. Trends Biotechnol. 27, 644651.
Bachman, P., 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.
22 (6), 12071222.
Barnard, A.-C., Nijhof, A., Fick, W., Stutzer, C., Maritz-Olivier, C., 2012. RNAi in arthro-
pods: insight into the machinery and applications for understanding the pathogen-vector
interface. Genes 3 (4), 702741.
Bartel, D.P., 2009. MicroRNAs: target recognition and regulatory functions. Cell 136 (2),
215233.
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 cole-
opteran insect pests through RNA interference. Nat. Biotechnol. 25 (11), 13221326.
Baum, J.A., CaJacob, C.A., Feldmann, P., Heck, G.R., Nooren, I., Plaetinck, G.,
Maddelein, W., Vaughn, T., 2011. Methods for control of insect infestation in plants
and compositions thereof. US Patent No. 7,943,819.
Baum, J.A., Evdokimov, A.G., Moshiri, F., Rydel, T.J., Sturman, E.J., von Rechenberg, M.,
Vu, H., Wollacott, A.M., Zheng, M., 2013. Proteins toxic to hemipteran insect species.
US Patent Application No. 20130269060 A1.
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 (1),
3846.
Belles, X., 2010. Beyond Drosophila, RNAi in vivo and functional genomics in insects. Annu.
Rev. Entomol. 55, 111128.
Bolognesi, R., Ramaseshadri, P., Anderson, J., Bachman, P., Clinton, W., Flannagan, R.,
Ilagan, O., Lawrence, C., Levine, S., Moar, W., Mueller, G., 2012. Characterizing
the mechanism of action of double-stranded RNA activity against western corn
rootworm Diabrotica virgifera virgifera LeConte. PLoS One 7 (10), e47534.
Boutros, M., Kiger, A.A., Armknecht, S., Kerr, K., Hild, M., Koch, B., Haas, S.A., Paro, R.,
Perrimon, N., 2004. Genome-wide RNAi analysis of growth and viability in Drosophila
cells. Science 303 (5659), 832835.
Boutros, M., Ahringer, J., 2008. The art and design of genetic screens: RNA interference.
Nat. Rev. Genet. 9 (7), 554566.
Burand, J.P., Hunter, W.B., 2013. RNAi, future in insect management. J. Invertebr. Pathol.
112 (Suppl. 1), S68S74.
Campbell, E.M., Budge, G.E., Bowman, A.S., 2010. Gene-knockdown in the honey bee
mite Varroa destructor by a non-invasive approach: studies on a glutathione
S-transferase. Parasit. Vectors 3, 73.
Chen, J., Zhang, D., Yao, Q., Zhang, J., Dong, X., Tian, H., Zhang, W., 2010. Feeding-
based RNA interference of a trehalose phosphate synthase gene in the brown plan-
thopper, Nilaparvata lugens. Insect Mol. Biol. 19 (6), 777786.
Christiaens, O., Sweveres, L., Smagghe, G., 2014. DsRNA degradation in the pea aphid
(Acyrthosiphon pisum) associated with lack of response in RNAi feeding and injection
assay. Peptides 53, 307314. http://dx.doi.org/10.1016/j.peptides.2013.12.014.
RNAi-Mediated Insect Pest Management 287
Clemens, J.C., Worby, C.A., Simonson-Leff, N., Muda, M., Maehama, T., Hemmings, B.A.,
Dixon, J.E., 2000. Use of double-stranded RNA interference in Drosophila cell lines to
dissect signal transduction pathways. Proc. Natl. Acad. Sci. U. S. A. 97 (12), 64996503.
Codex, 2003. Guideline for the conduct of food safety assessment of foods derived from
recombinant-DNA plants (CAC/GL 45-2003). Monsanto Technical, Rome, Italy,
pp. 726.
Cottrill, K.A., Chan, S.Y., 2014. Diet-derived microRNAs: separating the dream from real-
ity. microRNA Diagn. Ther. 1, 4657.
Coy, M.R., Sanscrainte, N.D., Chalaire, K.C., Inberg, A., Maayan, I., Glick, E., Paldi, N.,
Becnel, J.J., 2012. Gene silencing in adult Aedes aegypti mosquitoes through oral delivery
of double-stranded RNA. J. Appl. Entomol. 136 (10), 741748.
Dickinson, B., Zhang, Y., Petrick, J.S., Heck, G., Ivashuta, S., Marshall, W.S., 2013. Lack
of detectable oral bioavailability of plant microRNAs after feeding in mice. Nat. Bio-
technol. 31 (11), 965967.
Dietzl, G., Chen, D., Schnorrer, F., Su, K.C., Barinova, Y., Fellner, M., Gasser, B.,
Kinsey, K., Oppel, S., Scheiblauer, S., Couto, A., Marra, V., Keleman, K., Dickson, B.J.,
2007. A genome-wide transgenic RNAi library for conditional gene inactivation in Dro-
sophila. Nature 448, 151156.
Djebali, S., Davis, C.A., Merkel, A., Dobin, A., Lassmann, T., et al., 2012. Landscape of
transcription in human cells. Nature 489, 101108. http://dx.doi.org/10.1038/
nature11233.
Dow, J., 1992. pH gradients in lepidopteran midgut. J. Exp. Biol. 172 (1), 355375.
EFSA, 2006. Opinion of the Scientific Panel on Genetically Modified Organisms on a request
from the Commission related to the notification (Reference C/SE/96/3501) for the
placing on the market of genetically modified potato EH92-527-1 with altered starch
composition, for cultivation and production of starch, under Part C of Directive
2001/18/EC from BASF Plant Science. EFSA J. 323, 120.
Elbashir, S.M., Lendeckel, W., Tuschl, T., 2001. RNA interference is mediated by 21- and
22-nucleotide RNAs. Genes Dev. 15 (2), 188200.
FDA, 1994. Biotechnology of food. Monsanto technical report BG 94-4.
Feinberg, E.H., Hunter, C.P., 2003. Transport of dsRNA into cells by the transmembrane
protein SID-1. Science 301, 15451547.
Figueira-Mansur, J., Ferreira-Pereira, A., Mansur, J.F., Franco, T.A., Alvarenga, E.S.L.,
Sorgine, M.H.F., Neves, B.C., Melo, A.C.A., Leal, W.S., Masuda, H., Moreira, M.F.,
2013. Silencing of P-glycoprotein increases mortality in temephos-treated Aedes aegypti
larvae. Insect Mol. Biol. 22 (6), 648658.
Fire, A., Xu, S., Montgomery, M.K., Kostas, S.A., Driver, S.E., Mello, C.C., 1998. Potent
and specific genetic interference by double-stranded RNA in Caenorhabditis elegans.
Nature 391, 806811.
Furusawa, T., Takayama, E., Ishihara, R., Hayashi, Y., 1993. Double-stranded ribonuclease
activity in the digestive juice and midgut of the silkworm, Bombyx mori. Comp. Bio-
chem. Physiol. A Mol. Integr. Physiol. 104, 795801.
Garbian, Y., Maori, E., Kalev, H., Shafir, S., Sela, I., 2012. Bidirectional transfer of RNAi
between honey bee and Varroa destructor: Varroa gene silencing reduces varroa popula-
tion. PLoS Pathog. 8 (12), e1003035.
Garbutt, J.S., 2011. RNA Interference in Insects, Persistence and Uptake of Double-
Stranded RNA and Activation of RNAi Genes. (Ph.D. thesis). University of Bath.
Garbutt, J.S., Belles, X., Richards, E.H., Reynolds, S.E., 2013. Persistence of double-stranded
RNA in insect hemolymph as a potential determiner of RNA interference success,
Evidence from Manduca sexta and Blattella germanica. J. Insect Physiol. 59 (2), 171178.
Gassmann, A.J., Onstad, D.W., Pittendrigh, B.R., 2009. Evolutionary analysis of herbivo-
rous insects in natural and agricultural environments. Pest Manag. Sci. 65 (11),
11741181.
288 James A. Baum and James K. Roberts
Georghiou, G.P., Lagunes-Tejeda, A., Food and Agriculture Organization of the United
Nations, 1991. The Occurrence of Resistance to Pesticides in Arthropods: An Index
of Cases Reported Through 1989. Georghiou, G.P., Lagunes-Tejeda, A., FAO, Rome.
Gong, L., Yang, X., Zhang, B., Zhong, G., Hu, M., 2011. Silencing of Rieske iron-sulfur
protein using chemically synthesised siRNA as a potential biopesticide against Plutella
xylostella. Pest Manag. Sci. 67 (5), 514520.
Gong, L., Chen, Y., Hu, Z., Hu, M., 2013. Testing insecticidal activity of novel chemically
synthesized sirna against Plutella xylostella under laboratory and field conditions. PLoS
One 8 (5), e62990.
Griebler, M., Westerlund, S.A., Hoffmann, K.H., Meyering-Vos, M., 2008. RNA interfer-
ence with the allatoregulating neuropeptide genes from the fall armyworm Spodoptera
frugiperda and its effects on the JH titer in the hemolymph. J. Insect Physiol. 54 (6),
9971007.
Gu, L., Knipple, D.C., 2013. Recent advances in RNA interference research in insects:
Implications for future insect pest management strategies. Crop. Prot. 45, 3640.
Gu, J., Liu, M., Deng, Y., Peng, H., Chen, X., 2011. Development of an efficient recom-
binant mosquito densovirus-mediated RNA interference system and its preliminary
application in mosquito control. PLoS One 6 (6), e21329.
Hannon, G.J., 2002. RNA interference. Nature 418, 244251.
He, B., Chu, Y., Yin, M., Mullen, K., An, C., Shen, J., 2013. Fluorescent nanoparticle deliv-
ered dsRNA toward genetic control of insect pests. Adv. Mater. 25 (33), 45804584.
Hegedus, D., Erlandson, M., Gillott, C., Toprak, U., 2009. New insights into peritrophic
matrix synthesis, architecture, and function. Annu. Rev. Entomol. 54 (1), 285302.
Heisel, S.E., Zhang, Y., Allen, E., Guo, L., Reynolds, T.L., Yang, X., Kovalic, D.,
Roberts, J.K., 2008. Characterization of unique small RNA populations from rice grain.
PLoS One 3 (8), e2871.
Hinas, A., Wright, A.J., Hunter, C.P., 2012. SID-5 Is an endosome-associated protein
required for efficient systemic RNAi in C. elegans. Curr. Biol. 22 (20), 19381943.
Honeybee Genome Sequencing Consortium, 2006. Insights into social insects from the
genome of the honeybee Apis mellifera. Nature 443, 931949.
Houck, J.C., Berman, L.B., 1958. Serum ribonuclease activity. J. Appl. Physiol. 12, 473476.
Huang, F., Andow, D.A., Buschman, L.L., 2011. Success of the high-dose/refuge resistance
management strategy after 15 years of Bt crop use in North America. Entomol. Exp.
Appl. 140 (1), 116.
Hunter, W., Ellis, J., van Engelsdorp, D., Hayes, J., Westervelt, D., Glick, E., Williams, M.,
Sela, I., Maori, E., Pettis, J., Cox-Foster, D., Paldi, N., 2010. Large-scale field application
of RNAi technology reducing Israeli acute paralysis virus disease in honey bees (Apis
mellifera, Hymenoptera: Apidae). PLoS Pathog. 6 (12), e1001160.
Huvenne, H., Smagghe, G., 2010. Mechanisms of dsRNA uptake in insects and potential of
RNAi for pest control: a review. J. Insect Physiol. 56 (3), 227235.
ILSI-CERA, 2011. Problem Formulation for the Environmental Risk Assessment of RNAi
Plants. Center for Environmental Risk Assessment, International Life Sciences Institute
Research Foundation, Washington, D.C.
International Silkworm Genome Consortium, 2008. The genome of a lepidopteran model
insect, the silkworm Bombyx mori. Insect Biochem. Mol. Biol. 38, 10361045.
Ivashuta, S.I., Petrick, J.S., Heisel, S.E., Zhang, Y., Guo, L., Reynolds, T.L., Rice, J.F.,
Allen, E., Roberts, J.K., 2009. Endogenous small RNAs in grain: semi-quantification
and sequence homology to human and animal genes. Food Chem. Toxicol. 47 (2),
353360.
Jain, K.K., 2008. Stability and delivery of RNA via the gastrointestinal tract. Curr. Drug
Deliv. 5, 2731.
RNAi-Mediated Insect Pest Management 289
James, C., 2013. Global status of commercialized biotech/GM crops: 2013. ISAAA brief no.
46, ISAAA, Ithaca, NY.
Jensen, P.D., Zhang, Y., Wiggins, B.E., Petrick, J.S., Zhu, J., Kerstetter, R.A., Heck, G.R.,
Ivashuta, S.I., 2013. Computational sequence analysis of predicted long dsRNA trans-
criptomes of major crops reveals sequence complementarity with human genes. GM
Crops and Food: Biotechnology in Agriculture and the Food Chain 4, 9097.
Jose, A.M., Smith, J.J., Hunter, C.P., 2009. Export of RNA silencing from C. elegans tissues
does not require the RNA channel SID-1. Proc. Natl. Acad. Sci. U. S. A. 106 (7),
22832288.
Jose, A.M., Kim, Y.A., Leal-Ekman, S., Hunter, C.P., 2012. Conserved tyrosine kinase pro-
motes the import of silencing RNA into Caenorhabditis elegans cells. Proc. Natl. Acad. Sci.
U. S. A. 109 (36), 1452014525.
Kennerdell, J.R., Carthew, R.W., 1998. Use of dsRNA-mediated genetic interference
to demonstrate that frizzled and frizzled 2 act in the wingless pathway. Cell
95, 10171026.
Khajuria, C., Buschman, L.L., Chen, M.-S., Muthukrishnan, S., Zhu, K.Y., 2010. A gut-
specific chitinase gene essential for regulation of chitin content of peritrophic matrix
and growth of Ostrinia nubilalis larvae. Insect Biochem. Mol. Biol. 40 (8), 621629.
Khajuria, C., Li, H., Narva, K., Rangasamy, M., Siegfried, B., 2013. Effectiveness of dsRNA
versus siRNA in RNAi mediated gene knock-down in western corn rootworm
(Diabrotica virgifera virgifera). In: Program and Abstracts, 46th Annual Meeting of the Soci-
ety for Invertebrate Pathology Conference on Invertebrate Pathology and Microbial
Control. Society for Invertebrate Pathology. Pittsburgh, PA.
Koci, J., Ramaseshadri, P., Bolognesi, R., Segers, G., Flannagan, R., Park, Y., 2014. Ultra-
structural changes caused by Snf7 RNAi in larval enterocytes of western corn rootworm
Diabrotica virgifera virgifera Le Conte. PLoS One 9 (1), e83985.
Koizumi, K., Higashida, H., Yoo, S., Islam, M.S., Ivanov, A.I., Guo, V., Pozzi, P., Yu, S.H.,
Rovescalli, A.C., Tang, D., Nirenberg, M., 2007. RNA interference screen to identify
genes required for Drosophila embryonic nervous system development. Proc. Natl. Acad.
Sci. U. S. A. 104, 56265631.
Kontogiannatos, D., Swevers, L., Maenaka, K., Park, E.Y., Iatrou, K., Kourti, A., 2013.
Functional characterization of a juvenile hormone esterase related gene in the moth
Sesamia nonagrioides through RNA interference. PLoS One 8 (9), e73834.
Kumar, M., Gupta, G.P., Rajam, M.V., 2009. Silencing of acetylcholinesterase gene of Hel-
icoverpa armigera by siRNA affects larval growth and its life cycle. J. Insect Physiol. 55 (3),
273278.
Kumar, A., Wang, S., Ou, R., Samrakandi, M., Beerntsen, B.T., Sayre, R.T., 2013. Devel-
opment of an RNAi based microalgal larvicide to control mosquitoes. MalariaWorld J.
4, 6.
Kupferschmidt, K., 2013. A lethal dose of RNA. Science 341, 732733.
Kwon, D.H., Park, J.H., Lee, S.H., 2013. Screening of lethal genes for feeding RNAi by leaf
disc-mediated systematic delivery of dsRNA in Tetranychus urticae. Pest. Biochem. Phy-
siol. 105 (1), 6975.
Lee, J.T., 2012. Epigenetic regulation by long noncoding RNAs. Science 338, 14351439.
Lehane, M.J., 1997. Peritrophic matrix structure and function. Annu. Rev. Entomol.
42, 525550.
Li, J., Chen, Q., Lin, Y., Jiang, T., Wu, G., Hua, H., 2011a. RNA interference in Nilaparvata
lugens (Homoptera, Delphacidae) based on dsRNA ingestion. Pest Manag. Sci. 67 (7),
852859.
Li, X., Zhang, M., Zhang, H., 2011b. RNA interference of four genes in adult Bactrocera
dorsalis by feeding their dsRNAs. PLoS One 6 (3), e17788.
290 James A. Baum and James K. Roberts
Li, H., Jiang, W., Zhang, Z., Xing, Y., Li, F., 2013a. Transcriptome analysis and screening
for potential target genes for RNAi-mediated pest control of the beet armyworm,
Spodoptera exigua. PLoS One 8 (6), e65931.
Li, J., Wang, X.-P., Wang, M.-Q., Ma, W.-H., Hua, H.-X., 2013b. Advances in the use of
the RNA interference technique in Hemiptera. Insect Sci. 20 (1), 3139.
Liu, J., Smagghe, G., Swevers, L., 2013. Transcriptional response of BmToll9-1 and RNAi
machinery genes to exogenous dsRNA in the midgut of Bombyx mori. J. Insect Physiol.
59 (6), 646654.
Loretz, B., Foger, F., Werle, M., Bernkop-Schnurch, A., 2006. Oral gene delivery: strategies
to improve stability of pDNA towards intestinal digestion. J. Drug Target. 14, 311319.
Luo, Y., Wang, X., Yu, D., Chen, B., Kang, L., 2013. Differential responses of migratory
locusts to systemic RNA interference via double-stranded RNA injection and feeding.
Insect Mol. Biol. 22 (5), 574583.
Manzano-Roman, R., Oleaga, A., Perez-Sanchez, R., Siles-Lucas, M., Rollinson, D.,
Hay, S.I., 2012. Gene silencing in parasites: current status and future prospects. Adv.
Parasitol. 78, 155.
Mao, J., Zeng, F., 2012. Feeding-based RNA interference of a gap gene is lethal to the pea
aphid, Acyrthosiphon pisum. PLoS One 7 (11), e48718.
Mao, J., Zeng, F., 2014. Plant-mediated RNAi of a gap gene-enhanced tobacco tolerance
against the Myzus persicae. Transgenic Res. 23 (1), 145152.
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 (11),
13071313.
Mao, Y.-B., Tao, X.-Y., Xue, X.-Y., Wang, L.-J., Chen, X.-Y., 2011. Cotton plants
expressing CYP6AE14 double-stranded RNA show enhanced resistance to bollworms.
Transgenic Res. 20 (3), 665673.
Mao, Y.-B., Xue, X.-Y., Tao, X.-Y., Yang, C.-Q., Wang, L.-J., Chen, X.-Y., 2013. Cys-
teine protease enhances plant-mediated bollworm RNA interference. Plant Mol. Biol.
83 (12), 119129.
McDougall, P., 2011. The cost and time involved in the discovery, development and autho-
risation of a new plant biotechnology derived trait, a consultancy study for Crop Life
International. http://www.croplife.org/PhillipsMcDougallStudy.
McEwan, D.L., Weisman, A.S., Hunter, C.P., 2012. Uptake of extracellular double-stranded
RNA by SID-2. Mol. Cell 47 (5), 746754.
Meyering-Vos, M., Muller, A., 2007. RNA interference suggests sulfakinins as satiety effec-
tors in the cricket Gryllus bimaculatus. J. Insect Physiol. 53 (8), 840848.
Miller, S.C., Brown, S.J., Tomoyasu, Y., 2008. Larval RNAi in Drosophila? Dev. Genes
Evol. 218, 505510.
Miller, S.C., Miyata, K., Brown, S.J., Tomoyasu, Y., 2012. Dissecting systemic RNA inter-
ference in the red flour beetle Tribolium castaneum, parameters affecting the efficiency of
RNAi. PLoS One 7 (10), e47431.
Mutti, N.S., Park, Y., Reese, J.C., Reeck, G.R., 2006. RNAi knockdown of a salivary tran-
script leading to lethality in the pea aphid, Acyrthosiphon pisum. J. Insect Sci. 6, 38.
Mutti, N.S., Louis, J., Pappan, L.K., Pappan, K., Begum, K., Chen, M.-S., Park, Y.,
Dittmer, N., Marshall, J., Reese, J.C., Reeck, G.R., 2008. A protein from the salivary
glands of the pea aphid, Acyrthosiphon pisum, is essential in feeding on a host plant. Proc.
Natl. Acad. Sci. U. S. A. 105 (29), 99659969.
Nayak, A., Tassetto, M., Kunitomi, M., Andino, R., 2013. RNA interference-mediated
intrinsic antiviral immunity in invertebrates. In: Cullen, B.R. (Ed.), Intrinsic Immunity,
Current Topics in Microbiology and Immunology. In: 371, Springer-Verlag, Berlin,
Heidelberg, pp. 183200.
RNAi-Mediated Insect Pest Management 291
Nuez, I., Felix, M.-A., 2012. Evolution of susceptibility to ingested double-stranded RNAs
in Caenorhabditis nematodes. PLoS One 7 (1), e29811.
Nunes, F.M.F., Simoes, Z.L.P., 2009. A non-invasive method for silencing gene transcrip-
tion in honeybees maintained under natural conditions. Insect Biochem. Mol. Biol.
39 (2), 157160.
ONeill, M.J., Bourre, L., Melgar, S., ODriscoll, C.M., 2011. Intestinal delivery of non-viral
gene therapeutics: physiological barriers and preclinical models. Drug Discov. Today
16, 203218.
Paim, R.M.M., Araujo, R.N., Lehane, M.J., Gontijo, N.F., Pereira, M.H., 2013. Applica-
tion of RNA interference in triatomine (Hemiptera, Reduviidae) studies. Insect Sci.
20 (1), 4052.
Park, N.J., Li, Y., Yu, T., Brinkman, B.M., Wong, D.T., 2006. Characterization of RNA in
saliva. Clin. Chem. 52, 988994.
Parrott, W., Chassy, B., Ligon, J., Meyer, L., Petrick, J., Zhou, J., Herman, R., Delaney, B.,
Levine, M., 2010. Application of food and feed safety assessment principles to evaluate
transgenic approaches to gene modulation in crops. Food Chem. Toxicol.
48, 17731790.
Petrick, J.S., Brower-Toland, B., Jackson, A.L., Kier, L.D., 2013. Safety assessment of food
and feed from biotechnology-derived crops employing RNA-mediated gene regulation
to achieve desired traits: a scientific review. Regul. Toxicol. Pharmacol. 66, 167176.
Pilot, F., Philippe, J.M., Lemmers, C., Chauvin, J.P., Lecuit, T., 2006. Developmental con-
trol of nuclear morphogenesis and anchoring by charleston, identified in a functional
genomic screen of Drosophila cellularisation. Development 133, 711723.
Pitino, M., Coleman, A.D., Maffei, M.E., Ridout, C.J., Hogenhout, S.A., 2011. Silencing of
aphid genes by dsRNA feeding from plants. PLoS One 6 (10), e25709.
Prado, J.R., Segers, G., Voelker, T., Carson, D., Dobert, R., Phillips, J., Cook, K.,
Cornejo, C., Monken, J., Grapes, L., Reynolds, T., Martino-Catt, S., 2014. Genetically
engineered crops: from idea to product. Annu. Rev. Plant Biol. 65, 769790.
Pridgeon, J.W., Zhao, L., Becnel, J.J., Strickman, D.A., Clark, G.G., Linthicum, K.J., 2008.
Topically applied AaeIAP1 double-stranded RNA kills female adults of Aedes aegypti.
J. Med. Entomol. 45, 414420.
Rajagopal, R., Sivakumar, S., Agrawal, N., Malhotra, P., Bhatnagar, R.K., 2002. Silencing
of midgut aminopeptidase N of Spodoptera litura by double-stranded RNA establishes its
role as Bacillus thuringiensis toxin receptor. J. Biol. Chem. 277 (49), 4684946851.
Ramaseshadri, P., Segers, G., Flannagan, R., Wiggins, E., Clinton, W., Ilagan, O.,
McNulty, B., Clark, T., Bolognesi, R., 2013. Physiological and cellular responses caused
by RNAi- mediated suppression of Snf7 orthologue in western corn rootworm Diabrotica
virgifera virgifera larvae. PLoS One 8 (1), e54270.
Rangasamy, M., Siegfried, B.D., 2012. Validation of RNA interference in western corn
rootworm Diabrotica virgifera virgifera LeConte (Coleoptera, Chrysomelidae) adults. Pest
Manag. Sci. 68 (4), 587591.
Redenbaugh, K., Hiatt, W., Martineau, B., Kramer, M., Sheehy, R., Sanders, R.,
Houck, C., Emlay, D., 1992. Safety Assessment of Genetically Engineered Fruits and
Vegetables: A Case Study of the Flavr SavrTM Tomato. CRC Press, Inc., Boca
Raton, FL
Rodrguez-Cabrera, L., Trujillo-Bacallao, D., Borras-Hidalgo, O., Wright, D.J., Ayra-
Pardo, C., 2010. RNAi-mediated knockdown of a Spodoptera frugiperda trypsin-like
serine-protease gene reduces susceptibility to a Bacillus thuringiensis Cry1Ca1 protoxin.
Environ. Microbiol. 12 (11), 28942903.
Roignant, J.Y., Carre, C., Mugat, B., Szymczak, D., Lepesant, J.A., Antoniewski, C., 2003.
Absence of transitive and systemic pathways allows cell-specific and isoform-specific
RNAi in Drosophila. RNA 9 (3), 299308.
292 James A. Baum and James K. Roberts
Roush, R.T., 1998. Two-toxin strategies for management of insecticidal transgenic crops:
can pyramiding succeed where pesticide mixtures have not? Philos. Trans. R. Soc. Lond.
Ser. B Biol. Sci. 353 (1376), 17771786.
Saleh, M.-C., van Rij, R.P., Hekele, A., Gillis, A., Foley, E., OFarrell, P.H., Andino, R.,
2006. The endocytic pathway mediates cell entry of dsRNA to induce RNAi silencing.
Nat. Cell Biol. 8 (8), 793802.
Saleh, M.C., Tassetto, M., van Rij, R.P., Goic, B., Gausson, V., Berry, B., Jacquier, C.,
Antoniewski, C., Andino, R., 2009. Antiviral immunity in Drosophila requires systemic
RNA interference spread. Nature 458 (7236), 346350.
Shakesby, A.J., Wallace, I.S., Isaacs, H.V., Pritchard, J., Roberts, D.M., Douglas, A.E., 2009.
A water-specific aquaporin involved in aphid osmoregulation. Insect Biochem. Mol.
Biol. 39 (1), 110.
Shih, J.D., Hunter, C.P., 2011. SID-1 is a dsRNA-selective dsRNA-gated channel. RNA
17, 10571065.
Shih, J.D., Fitzgerald, M.C., Sutherlin, M., Hunter, C.P., 2009. The SID-1 double-stranded
RNA transporter is not selective for dsRNA length. RNA 15, 384390.
Silva, C.P., Silva, J.R., Vasconcelos, F.F., Petretski, M.D.A., DaMatta, R.A., Ribeiro, A.F.,
Terra, W.R., 2004. Occurrence of midgut perimicrovillar membranes in paraneopteran
insect orders with comments on their function and evolutionary significance. Arthropod
Struct. Dev. 33 (2), 139148.
Singh, A.D., Wong, S., Ryan, C.P., Whyard, S., 2013. Oral delivery of double-stranded
RNA in larvae of the yellow fever mosquito, Aedes aegypti: implications for pest mosquito
control. J. Insect Sci. 13, 69.
Snow, J.W., Hale, A.E., Isaacs, S.K., Baggish, A.L., Chan, S.Y., 2013. Ineffective delivery of
diet-derived microRNAs to recipient animal organisms. RNA Biol. 10 (7), 11071116.
Stevens, C.E., Hume, I.D., 2004. Comparative Physiology of the Vertebrate Digestive Sys-
tem, second ed. Cambridge University Press, Cambridge.
Storer, N.P., Babcock, J.M., Schlenz, M., Meade, T., Thompson, G.D., Bing, J.W.,
Huckaba, R.M., 2010. Discovery and characterization of field resistance to Bt maize:
Spodoptera frugiperda (Lepidoptera: Noctuidae) in Puerto Rico. J. Econ. Entomol.
103 (4), 10311038.
Surakasi, V.P., Mohamed, A.A.M., Kim, Y., 2011. RNA interference of 1 integrin subunit
impairs development and immune responses of the beet armyworm, Spodoptera exigua.
J. Insect Physiol. 57 (11), 15371544.
Swevers, L., Vanden Broeck, J., Smagghe, G., 2013. The possible impact of persistent virus
infection on the function of the RNAi machinery in insects: a hypothesis. Front. Physiol.
4http://dx.doi.org/10.3389/fphys.2013.00319.
Tabashnik, B.E., 2008. Delaying insect resistance to transgenic crops. Proc. Natl. Acad. Sci.
U. S. A. 105 (49), 1902919030.
Tang, T., Zhao, C., Feng, X., Liu, X., Qiu, L., 2012. Knockdown of several components of
cytochrome P450 enzyme systems by RNA interference enhances the susceptibility of
Helicoverpa armigera to fenvalerate. Pest Manag. Sci. 68 (11), 15011511.
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., Felfoldi, G., Fujiwara, H., Futahashi, R.,
Gandhe, A.S., Gatehouse, H.S., Gatehouse, L.N., Giebultowicz, J.M., Gomez, I.,
Grimmelikhuijzen, C.J., Groot, A.T., Hauser, F., Heckel, D.G., Hegedus, D.D.,
Hrycaj, S., Huang, L., Hull, J.J., Iatrou, K., Iga, M., Kanost, M.R., Kotwica, J.,
Li, C., Li, J., Liu, J., Lundmark, M., Matsumoto, S., Meyering-Vos, M., Millichap, P.J.,
Monteiro, 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.,
RNAi-Mediated Insect Pest Management 293
Soberon, M., Strand, M.R., Tomita, S., 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 (2),
231245.
The International Aphid Genomics Consortium, 2010. Genome sequence of the pea aphid
Acyrthosiphon pisum. PLoS Biol. 8 (2), e1000313.
Tian, H., Peng, H., Yao, Q., Chen, H., Xie, Q., Tang, B., Zhang, W., 2009. Developmental
control of a lepidopteran pest Spodoptera exigua by ingestion of bacteria expressing
dsRNA of a non-midgut gene. PLoS One 4 (7), e6225.
Tijsterman, M., May, R.C., Simmer, F., Okihara, K.L., Plasterk, R.H.A., 2004. Genes
required for systemic RNA interference in Caenorhabditis elegans. Curr. Biol. 14, 111116.
Timmons, L., Fire, A., 1998. Specific interference by ingested dsRNA. Nature 395, 854.
Timmons, L., Court, D.L., Fire, A., 2001. Ingestion of bacterially expressed dsRNAs can
produce specific and potent genetic interference in Caenorhabditis elegans. Gene
263, 103112.
Tomari, Y., Zamore, P.D., 2005. Perspective, machines for RNAi. Genes Dev. 19 (5),
517529.
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 (1), R10.
Tribolium Genome Sequence Consortium, 2008. The genome of the model beetle and pest
Tribolium castaneum. Nature 452, 949955.
Turner, C.T., Davy, M.W., MacDiarmid, R.M., Plummer, K.M., Birch, N.P.,
Newcomb, R.D., 2006. RNA interference in the light brown apple moth, Epiphyas
postvittana Walker induced by double-stranded RNA feeding. Insect Mol. Biol.
15 (3), 383391.
Ulvila, J., Parikka, M., Kleino, A., Sormunen, R., Ezekowitz, R.A., Kocks, C., Ramet, M.,
2006. Double-stranded RNA is internalized by scavenger receptor-mediated endocyto-
sis in Drosophila S2 cells. J. Biol. Chem. 281 (20), 1437014375.
Upadhyay, S.K., Chandrashekar, K., Thakur, N., Verma, P.C., Borgio, J.F., Singh, P.K.,
Tuli, R., 2011. RNA interference for the control of whiteflies (Bemisia tabaci) by oral
route. J. Biosci. 36, 153161.
Vachon, V., Laprade, R., Schwartz, J.-L., 2012. Current models of the mode of action of
Bacillus thuringiensis insecticidal crystal proteins: a critical review. J. Invertebr. Pathol.
111, 112.
Vandenborre, G., Van Damme, E.J.M., Smagghe, G., 2009. Natural products: plant lectins as
important tools in controlling pest insects. In: Ishaaya, I., Horowitz, A.R. (Eds.), Bio-
rational Control of Arthropod Pests. Springer, Netherlands, pp. 163187.
Vander Meer, R.K., Choi, M.Y., 2013. Formicidae (ant) control using double-stranded
RNA constructs. US Patent No. 8,575,328.
Varkouhi, A.K., Scholte, M., Storm, G., Haisma, H.J., 2011. Endosomal escape pathways for
delivery of biologicals. J. Control. Release 151 (3), 220228.
Walker, W.B., Allen, M.L., 2011. RNA interference-mediated knockdown of IAP in Lygus
lineolaris induces mortality in adult and pre-adult life stages. Entomol. Exp. Appl.
138, 8392.
Walshe, D.P., Lehane, S.M., Lehane, M.J., Haines, L.R., 2009. Prolonged gene knockdown
in the tsetse fly Glossina by feeding double stranded RNA. Insect Mol. Biol. 18 (1),
1119.
Wang, Y., Zhang, H., Li, H.-C., Miao, X.-X., 2011. Second-generation sequencing supply
an effective way to screen RNAi targets in large scale for potential application in pest
insect control. PLoS One 6 (4), e18644.
294 James A. Baum and James K. Roberts
Wang, J., Wu, M., Wang, B., Han, Z., 2013. Comparison of the RNA interference effects
triggered by dsRNA and siRNA in Tribolium castaneum. Pest Manag. Sci. 69, 781786.
Whangbo, J.S., Hunter, C.P., 2008. Environmental RNA interference. Trends Genet.
24 (6), 297305.
Wheeler, D., Darby, B.J., Todd, T.C., Herman, M.A., 2012. Several grassland soil nematode
species are insensitive to RNA-mediated interference. J. Nematol. 44, 92101.
Whyard, S., Singh, A.D., Wong, S., 2009. Ingested double-stranded RNAs can act as
species-specific insecticides. Insect Biochem. Mol. Biol. 39 (11), 824832.
Winston, W.M., Molodowitch, C., Hunter, C.P., 2002. Systemic RNAi in C. elegans
requires the putative transmembrane protein SID-1. Science 295, 24562459.
Winston, W.M., Sutherlin, M., Wright, A.J., Feinberg, E.H., Hunter, C.P., 2007.
Caenorhabditis elegans SID-2 is required for environmental RNA interference. Proc. Natl.
Acad. Sci. U. S. A. 104, 1056510570.
Witwer, K.W., Hirschi, K.D., 2014. Transfer and functional consequences of dietary micro-
RNAs invertebrates: concepts in search of corroboration. Bioessays 36http://dx.doi.org/
10.1002/bies.201300150.
Witwer, K.W., McAlexander, M.A., Queen, S.E., Adams, R.J., 2013. Real-time quantita-
tive PCR and droplet digital PCR for plant miRNAs in mammalian blood provide little
evidence for general uptake of dietary miRNAs. RNA Biol. 10 (7), 10801086.
Wuriyanghan, H., Rosa, C., Falk, B.W., 2011. Oral delivery of double-stranded RNAs and
siRNAs induces RNAi effects in the potato/tomato psyllid, Bactericerca cockerelli. PLoS
One 6 (11), e27736.
Wynant, N., Verlinden, H., Breugelmans, B., Simonet, G., Vanden Broeck, J., 2012. Tissue-
dependence and sensitivity of the systemic RNA interference response in the desert
locust, Schistocerca gregaria. Insect Biochem. Mol. Biol. 42 (12), 911917.
Wynant, N., Santos, D., Verdonck, R., Spit, J., Van Wielendaele, P., Vanden Broeck, J.,
2014. Identification, functional characterization and phylogenetic analysis of double
stranded RNA degrading enzymes present in the gut of the desert locust, Schistocerca
gregaria. Insect Biochem. Mol. Biol.. http://dx.doi.org/10.1016/j.ibmb.2013.12.008.
Xiong, Y., Zeng, H., Zhang, Y., Xu, D., Qiu, D., 2013. Silencing the HaHR3 gene by trans-
genic plant-mediated RNAi to disrupt Helicoverpa armigera development. Int. J. Biol. Sci.
9 (4), 370381.
Xu, W., Han, Z., 2008. Cloning and phylogenetic analysis of sid-1-like genes from aphids.
J. Insect Sci. 8, 30.
Xu, H.J., Chen, T., Ma, X.F., Xue, J., Pan, P.L., Zhang, X.C., Cheng, J.A., Zhang, C.X.,
2013. Genome-wide screening for components of small interfering RNA (siRNA) and
micro-RNA (miRNA) pathways in the brown planthopper, Nilaparvata lugens
(Hemiptera: Delphacidae). Insect Mol. Biol. 22 (6), 635647. http://dx.doi.org/
10.1111/imb.12051.
Yang, J., Han, Z.-j., 2014. Efficiency of different methods for dsRNA delivery in cotton
bollworm (Helicoverpa armigera). J. Integr. Agric. 13 (1), 115123.
Yao, J., Rotenberg, D., Afsharifar, A., Barandoc-Alviar, K., Whitfield, A.E., 2013. Devel-
opment of RNAi methods for Peregrinus maidis, the corn planthopper. PLoS One 8 (8),
e370243.
Yu, N., Christiaens, O., Liu, J., Niu, J., Cappelle, K., Caccia, S., Huvenne, H., Smagghe, G.,
2013. Delivery of dsRNA for RNAi in insects, an overview and future directions. Insect
Sci. 20, 414.
Zha, W., Peng, X., Chen, R., Du, B., Zhu, L., He, G., 2011. Knockdown of midgut genes
by dsRNA-transgenic plant-mediated RNA interference in the hemipteran insect
Nilaparvata lugens. PLoS One 6 (5), e20504.
RNAi-Mediated Insect Pest Management 295
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 (5), 683693.
Zhang, X., Zhang, J., Zhu, K., 2012a. Advances and prospects of RNAi technologies in
insect pest management. In: Liu, T., Kang, L. (Eds.), Recent Advances in Entomological
Research. Springer, Berlin Heidelberg, pp. 347358.
Zhang, Y., Wiggins, B.E., Lawrence, C., Petrick, J., Ivashuta, S., Heck, G., 2012b. Analysis
of plant-derived miRNAs in animal small RNA datasets. BMC Genomics 13, 381.
Zhang, Y.-l., Zhang, S.-z., Kulye, M., Wu, S.-r., Yu, N.-t., Wang, J.-h., Zeng, H.-m.,
Liu, Z.-x., 2012c. Silencing of molt-regulating transcription factor gene, CiHR3, affects
growth and development of sugarcane stem borer, Chilo infuscatellus. J. Insect Sci. 12 (91),
112.
Zhang, H., Li, H.-C., Miao, X.-X., 2013a. Feasibility, limitation and possible solutions of
RNAi-based technology for insect pest control. Insect Sci. 20, 1530.
Zhang, M., Zhou, Y., Wang, H., Jones, H.D., Gao, Q., Wang, D., Ma, Y., Xia, L., 2013b.
Identifying potential RNAi targets in grain aphid Sitobion avenae F. based on trans-
criptome profiling of its alimentary canal after feeding on wheat plants. BMC Genomics
14, 560.
Zhang, X., Liu, X., Ma, J., Zhao, J., 2013c. Silencing of cytochrome P450 CYP6B6 gene of
cotton bollworm (Helicoverpa armigera) by RNAi. Bull. Entomol. Res. 103 (5), 584591.
Zhao, L., Chen, J., 2013. Double stranded RNA constructs to control ants. US Patent Appli-
cation Publication No. 2013/0078212.
Zhao, Y., Yang, G., Wang-Pruski, G., You, M., 2008. Phyllotreta striolata (Coleoptera,
Chrysomelidae): arginine kinase cloning and RNAi-based pest control. Eur. J. Biochem.
105, 815822.
Zhou, X., Wheeler, M.M., Oi, F.M., Scharf, M.E., 2008. RNA interference in the termite
Reticulitermes flavipes through ingestion of double-stranded RNA. Insect Biochem. Mol.
Biol. 38 (8), 805815.
Zhou, J., Shum, K.-T., Burnett, J., Rossi, J., 2013. Nanoparticle-based delivery of RNAi
therapeutics, progress and challenges. Pharmaceuticals 61, 85107.
Zhu, F., Xu, J., Palli, R., Ferguson, J., Palli, S.R., 2011. Ingested RNA interference for man-
aging the populations of the Colorado potato beetle, Leptinotarsa decemlineata. Pest
Manag. Sci. 67 (2), 175182.
Zhu, J.-Q., Liu, S., Ma, Y., Zhang, J.-Q., Qi, H.-S., Wei, Z.-J., Yao, Q., Zhang, W.-Q.,
Li, S., 2012. Improvement of pest resistance in transgenic tobacco plants expressing
dsRNA of an insect-associated gene EcR. PLoS One 7 (6), e38572.
CHAPTER SIX
Contents
1. Introduction 298
2. Detection Methods and Current Status of Insect Resistance to Bt Crops 299
2.1 Definition of resistance 299
2.2 Resistance detection methods 301
2.3 Current status of field-evolved resistance to Bt crops 312
3. Resistance Mechanisms 314
3.1 Mode of action of Bt Cry toxins 314
3.2 Alterations in proteolytic processing of Cry toxins in resistant insects 318
3.3 Modifications of Cry toxin receptors in resistant insects 319
4. Genetic Diversity of Resistance and Implications for Resistance Management 324
4.1 Laboratory-selected and field-evolved resistance 324
4.2 Resistance dominance and the refuge strategy 328
4.3 Cross-resistance and the pyramid strategy 329
5. Conclusions 331
Acknowledgements 332
References 332
Abstract
Transgenic crops producing Bacillus thuringiensis Cry toxins (Bt crops) have been planted
globally to control some key pests. Benefited from implementation of proactive resis-
tance management strategies such as the refuge strategy and the pyramid strategy in
many countries, most of the target pests of Bt crops have been sustainably and effec-
tively controlled for nearly 20 years. However, several cases of field-evolved resistance to
Bt corn and Bt cotton have been documented, causing reduced field efficacy. Evolution
of resistance by target pests is a real threat to the continued success of Bt crops. It is
crucial to employ sensitive detection methods to monitor the evolution of the resis-
tance in the target insects and thereby adapt resistance management strategies
1. INTRODUCTION
Bacillus thuringiensis (Bt) is a gram-positive bacterium that is character-
ized by producing parasporal crystal proteins with insecticidal activity (Cry
toxins) during sporulation (Knowles, 1994; Schnepf et al., 1998). Bt sprays
have been widely used as a bioinsecticide in agriculture, forestry and mos-
quito control for several decades (Sanahuja et al., 2011). Transgenic crops
expressing genes encoding Bt toxins (Bt crops) were commercialized for
the first time in 1996 in the United States; since then, the cumulative total
of more than 570 million ha of Bt crops (mainly corn and cotton) has been
adopted globally ( James, 2013; Tabashnik et al., 2013). Intensive planting of
Bt crops inevitably creates strong selection pressure on the target insect pests,
thereby the resistance to these Bt crops evolved by target insect pests is con-
sidered a major threat to the durability of Bt crops (Carriere et al., 2010;
Gould, 1998; Huang et al., 2011; Tabashnik, 1994).
As a proactive effort to preserve the value and benefits of Bt crops, the
refuge strategy has been widely used to delay development of insect resis-
tance to Bt crops (Tabashnik et al., 2004). In theory, three key conditions
should be met in order to maximize the effectiveness of the refuge strategy:
recessive inheritance of resistance phenotype, low frequency of initial resis-
tance alleles and abundant susceptible insects provided by non-Bt host plants
nearby (Tabashnik et al., 2013). In practice, target insects have variable and
unpredictable inheritance patterns of the resistance trait and the initial fre-
quencies of resistance alleles within geographical populations. So, it is crucial
to employ sensitive detection methods to monitor the dynamic evolution of
the resistance alleles over space and time for the target insects and thereby
Insect Resistance to Bt Cry Toxins 299
to both single insects and whole populations, and the severity of resistance
can range from no effect on field control to field failure of an insecticide
application. Brent (1986) defined resistance as any heritable decrease in sen-
sitivity to a chemical within a pest population. Brent (1986) emphasized
that insecticide resistance can cause complete loss of action of an agrochem-
ical or may have little practical significance. Sawicki (1987) modified Crows
definition to Insecticide resistance marks a genetic change in response to
selection by toxicants that may impair control in the field. In the definitions
mentioned above, both a change in susceptibility to the insecticide detect-
able with bioassays in the laboratory and field control failure are considered
resistance. In other words, resistance to an insecticide need not result in loss
of insect pest control.
Tabashnik et al. (2009a, 2014) defined the term field-evolved
resistance as a genetically based decrease in susceptibility of a population
to a pesticide caused by exposure to the pesticide in the field. The term
field-evolved resistance applies to resistance in both pest and beneficial
organisms and does not necessarily imply loss of economic efficacy in the
field (Tabashnik et al., 2009a). This general definition favours proactive
detection and management of resistance.
Another definition of resistance proposed by the Insecticide Resistance
Action Committee (IRAC) is a heritable change in the sensitivity of a pest
population that is reflected in the repeated failure of a product to achieve the
expected level of control when used according to the label recommendation
for that pest species (http://www.irac-online.org/about/resistance). This
definition emphasizes a causal relationship between resistance and field con-
trol failure. Agrochemical industry intends to use this definition to validate
and confirm resistance as the cause of observed losses in field efficacy.
There are debates on the term, field-evolved resistance, in defining
insect resistance to Bt crops. A general definition of field-evolved
resistance considers that the primary goal of monitoring resistance to Bt
crops is to detect evolution of resistance early enough to enable proactive
management of resistant insects, and field control problems associated with
field-evolved resistance vary from none to severe (Tabashnik et al., 2008,
2013). A narrow definition of field-evolved resistance regards decreased
field efficacy and/or survival on Bt plants as the decisive criteria for defining
a case of field-evolved resistance (Moar et al., 2008; Sumerford et al., 2013).
To avoid this confusion, Tabashnik et al. (2014) defined practical
resistance as field-evolved resistance that reduces the efficacy of a pesticide
and has practical consequences for pest control.
Insect Resistance to Bt Cry Toxins 301
which kills all or nearly all individuals of the susceptible population, but few
or no resistant individuals. Resistance allele frequency can be readily esti-
mated if the genetic basis of resistance is known (Tabashnik et al., 2009a;
Zhang et al., 2012a). Compared with the concentrationresponse assay,
the diagnostic concentration test is more efficient for detecting an incipient
resistance outbreak (Roush and Miller, 1986).
When designing and implementing an efficient detection programme for
Bt resistance based on the concentrationresponse and/or diagnostic con-
centration assays, several key factors should be considered:
(1) Establishment of baseline susceptibility data and the calibration of a diag-
nostic concentration. It is important to survey the baseline susceptibility
of target pest populations sampled from geographical areas prior to wide
commercial adoption of a Bt crop. The baseline data can reveal the range
of geographical variations of unexposed field populations which will then
be necessary for defining susceptibility changes relating to exposure to Bt
crops. Similarly, it is necessary to calibrate the diagnostic concentrations
against a range both of unexposed field populations and of susceptible lab-
oratory strains (Forrester et al., 1993; Khakame et al., 2013; Wang et al.,
2010). Compared with the appropriate baseline comparators considering
natural susceptibility variations, a statistically significant increase in sur-
vival at a diagnostic concentration and/or in LC50 values or LC90 will
provide initial evidence of resistance evolution in the field.
(2) Maintaining a susceptible laboratory strain for comparison across years and
regions. For valid comparison with the susceptibility of field populations, it
is necessary to choose a reference strain to be kept in the laboratory, which
should be neither extremely susceptible nor unusually tolerant to the Bt
toxin concerned. Concurrent comparative susceptibility data for a specific
Bt toxin from the susceptible strain are essential for comparison with field
population susceptibility across years and regions.
(3) Sampling and testing. The scale of sampling depends on the aims and
types of resistance detection as well as on the mobility and population
structure of the target pest. Each collection needs to be representative of
the local field population. It is suggested that hundreds of insects should
be collected from many locations instead of thousands of insects from a
few locations (ffrench-Constant and Roush, 1990). Considering costs
of time and labour, it is often best that an extensive survey be done first,
followed by a more intensive monitoring of selected locations, which
have intensive adoption of Bt crops and possibly a higher risk of resis-
tance evolution.
Insect Resistance to Bt Cry Toxins 303
H. armigera, and one resistant strain of H. zea, both pests that are targeted
by Bt cotton, were ca. 10 times more resistant to Cry1Ac toxin than to
Cry1Ac protoxin. The SCD-r1 strain of H. armigera had ca. 500-fold
resistance to Cry1Ac toxin, but only 39-fold resistance to Cry1Ac
protoxin (Tabashnik et al., 2011; Xu et al., 2005). Similarly, the AR
strain of H. zea had more than 100-fold resistance to Cry1Ac toxin,
but only 10-fold to Cry1Ac protoxin (Anilkumar et al., 2008). For
Bt resistance monitoring of these two cotton pests, it could be better
to use Cry1Ac toxin for early detection of resistance.
2.2.2 F1 screen
The principles of complementation tests are as follows (Zhang et al., 2012a):
If two resistant strains are crossed, each with recessive alleles for resistance at
separate loci, allelic complementation will restore susceptibility (the wild-
type phenotype) in the progeny. However, if the recessive resistance alleles
occur at the same locus in different strains, the progeny from the cross
between strains will be resistant because they will inherit resistance alleles
at the same locus from both parents. Based on the principle of complemen-
tation tests, Gould et al. (1997) developed an F1 screen method to estimate
the frequency of Cry1Ac-resistance alleles in field populations of the tobacco
budworm, Heliothis virescens. Over 2000 single-pair families were made
among field-collected male moths and virgin female moths from the
laboratory-selected YHD2 strain of H. virescens, which has a high level of
Cry1Ac resistance conferred by a recessive gene (producing a truncated
cadherin). F1 offspring from each of 1025 families were screened with a dis-
criminating concentration of Cry1Ac that could distinguish heterozygous
from homozygous resistant individuals. Four of these families produced
3042% offspring with resistance to Cry1Ac, indicating that in each of
the four field-collected males was heterozygous and carried a recessive resis-
tance allele. By the F1 screen, the initial frequency of alleles for Cry1Ac resis-
tance was able to be estimated as 0.0015 in field populations of H. virescens
(Gould et al., 1997).
The F1 screen method has been used to determine initial Bt resistance
allele frequencies in field populations of several pests other than
H. virescens. A total of 286 single-pair crosses were made between field-
derived adults of the poplar leaf beetle, Chrysomela tremulae, and adults of
a laboratory strain with recessive resistance to Cry3Aa. F1 neonates from
each of the 176 single-pair families with enough F1 offspring were screened
with leaf discs from a Bt poplar line. Three of the 176 families screened
Insect Resistance to Bt Cry Toxins 305
Figure 6.1 F1 screen for resistance alleles in Helicoverpa armigera (Zhang et al., 2012a).
Field-caught male moths were crossed individually to homozygous female moths with a
recessive resistance allele at the cadherin locus (rcrc) from a laboratory strain of
H. armigera. F1 offspring from single-pair families were tested at a diagnostic concen-
tration of Cry1Ac. Expected survival of the F1 progeny in these bioassays depends on the
genotype of the field-caught male: 0% for a susceptible homozygote (ss), 50% for indi-
viduals with one susceptible allele and either any recessive resistance allele at the
cadherin locus (rc) or a dominant resistance allele at any locus (Rx) (genotypes rcs or
Rxs) and 100% for individuals with the genotypes rcrc or RxRx. To determine the domi-
nance of the resistance alleles detected with the F1 screen, survivors of the F1 screen are
crossed with a susceptible strain (ss) and the progeny tested at the diagnostic concen-
tration. Expected survival of the progeny of these F1 survivors ss crosses is 0% if the
resistance allele is recessive (progeny are rcs) and 50% if the allele is dominant (half of
progeny are Rxs and half are ss). It should be noted that F1 screen does not detect reces-
sive alleles at loci other than the same resistance locus as in the tester laboratory strain.
2.2.3 F2 screen
Andow and Alstad (1998) developed the F2 screen technique to detect reces-
sive alleles in field populations. The principle of an F2 screen is that F1 off-
spring of each mated female collected from the field are sib-mated to
produce F2 iso-female families. Each mated female carries four haploid
genomes, two from her own and two from her mate. If the field-collected
female or her mate carries one recessive resistance allele (r), the F2 offspring
of the iso-female family will have approximately 6.25% survival (rr) when
screened with a diagnostic concentration of Bt toxin or with Bt plants. Com-
pared with the diagnostic concentration assay, the F2 screen extends the sen-
sitivity to detect recessive resistance traits by more than an order of
magnitude. Compared with the F1 screen, the F2 screen can detect resistance
alleles at any locus rather than just the same locus where resistance alleles
occur in the laboratory strains (Andow and Alstad, 1998).
Because of its robustness, the F2 screen has been employed to detect
resistance allele frequency to Bt toxins for at least eight lepidopteran pests:
European corn borer, Ostrinia nubilalis (Andow et al., 1998, 2000;
Bourguet et al., 2003; Siegfried et al., 2014; Stodola et al., 2006), rice yellow
stem borer, Scirpophaga incertulas (Bentur et al., 2000), southwestern corn
borer, Diatraea grandiosella (Huang et al., 2007a), Mediterranean corn borer,
Sesamia nonagrioides (Andreadis et al., 2007), sugarcane borer, D. saccharalis
(Huang et al., 2007b, 2008, 2009), H. armigera (Liu et al., 2010; Mahon
et al., 2007, 2012; Xu et al., 2009; Zhang et al., 2012a), H. punctigera
(Downes et al. 2009, 2010a; Mahon et al. 2012) and H. virescens (Blanco
et al., 2009) and one coleopteran pest: C. tremulae (Genissel et al., 2003).
Because the F2 screen involves intensive input of time and resources, it
has been used only to determine initial resistance allele frequency for most
monitoring plans. However, F2 screen has been practically used as the main-
stay of the routine monitoring of Bt resistance in two cotton pests,
H. armigera and H. punctigera in Australia (Downes and Mahon, 2012a,b).
From 2002/2003 until 2010/2011, F2 screens were used to test 1222 iso-
female lines (4888 alleles) of H. armigera and 1558 (6232 alleles) iso-female
lines of H. punctigera for Cry1Ac resistance. Two H. armigera and three
H. punctigera lines were positive for a resistance allele to Cry1Ac. Based
on data pooled since 2002/2003, the estimated Cry1Ac resistance allele
frequency for H. armigera is 0.0006 (95% CI between 0.0001 and 0.002)
and for H. punctigera is 0.0006 (95% CI between 0.0002 and 0.001).
During the same period, F2 screens were used to test 1303 iso-female
lines (5212 alleles) of H. armigera and 1642 iso-female lines (6566 alleles)
308 Yidong Wu
that some resistance alleles are homozygous lethal if autozygous (as gen-
erated in F2 tests) but not as allozygous homozygotes (as generated in F1
tests). The hypothesis was extended to accommodate the possibility that
alleles at linked loci may be homozygous lethal. However, neither of two
tests of the hypothesis carried out to date provided evidence that any alleles
that confer resistance are associated with severe fitness costs (Mahon et al.,
2010), which does not support the hypothesis.
In China, diverse cadherin mutants associated with Cry1Ac resistance in
H. armigera were detected from F1 screens and F2 screens (Yang et al., 2007;
Zhang et al., 2012a; Zhao et al., 2010). A number of resistant strains with
autozygous genotypes of cadherin showed various resistance intensities to
Cry1Ac, ranging from 31- to 530-fold (Zhang et al., 2012a). Under a diag-
nostic concentration of Cry1Ac (1 g/cm2 diet surface), the AY148 strain
(r9r9) had 46% survival, the SCD-r1 strain (r1r1) had 92% survival and F1 off-
spring (r1r9) from AY148 SCD-r1 had 81% survival (Zhang et al., 2012a).
If one of the field parents of an iso-female line in an F2 screen carries a copy
of the r9 allele, the F2 offspring will have only 2.9% survival (6.25% 0.46)
under the diagnostic concentration, which will be scored negative. In con-
trast, if one of the field males carrying a copy of r9 allele (r9s) is crossed with a
female from SCD-r1 (r1r1) in an F1 screen, the F1 offspring (r1r9) will have
41% survival (81% 0.5) under the diagnostic concentration, which will be
scored positive. In F2 screens of Cry1Ac resistance alleles in two populations
of H. armigera from China, the parents of iso-female lines were scored as hav-
ing at least one resistance allele if F2 progeny survival was >3% instead of
>6.25% (Zhang et al., 2012a). Thus, Cry1Ac resistance allele frequencies
estimated by F1 screens were similar to that by F2 screens (Zhang et al.,
2012a). So, variability in resistance intensities among different autozygotes
(such as rara and rbrb) and/or allozygotes (rarb) might cause discrepancy in
results of F1 screens and F2 screens.
words, DNA screen can only pick up resistance data for which genetic basis
is already well characterized.
DNA-based screening has become applicable to Bt resistance since
cadherin mutations were confirmed to cause Cry1Ac resistance in three lep-
idopteran pests of cotton (Gahan et al., 2001; Morin et al., 2003; Xu et al.,
2005). In P. gossypiella, >8000 insects collected from cotton fields in
Arizona, California and Texas during 20012009 were screened with a
DNA-based diagnostic PCR for three resistance alleles from laboratory-
selected strains, and none of these three resistance alleles were detected
(Tabashnik et al., 2006, 2010). Similarly, the cadherin resistance allele from
a laboratory-selected strain of H. virescens was not detected in >7000 field-
collected individuals (Gahan et al., 2007). These results indicate extremely
low frequencies of the cadherin resistance alleles that could be detected, but
they do not exclude the presence of resistance alleles at either the cadherin
locus or other loci that could not be detected by the PCR methods used. In
both cases, however, bioassay data show that resistance remained rare in field
populations (Blanco et al., 2009; Tabashnik et al., 2010).
In H. armigera, an allele (r15) with a 55aa deletion in the intracellular
domain of cadherin (HaCad) was identified to confer non-recessive resis-
tance to Cry1Ac (Zhang et al., 2012b). Subsequently, a DNA-based
PCR method was developed to screen for the r15 allele in field populations
collected from the main cotton planting areas of China in 2011 and 2012
(Zhang et al., 2013). Three heterozygous r15 alleles were detected from
562 moths collected from northern China (with intensive Bt cotton plant-
ing), and the frequency of r15 allele was estimated to be 0.0027. However, no
r15 allele was detected from 314 moths collected from Xinjiang (with limited
Bt cotton use). Although all the r15 alleles have the same deletion in the
cDNA sequence, at least four different indels causing loss of exon 32 have
been detected in the genomic DNA sequences flanking exon 32 of HaCad.
Thus, designing a new method to detect the resistance alleles such as r15 at
cDNA level can avoid underestimating their frequency.
DNA screening is still in its infancy for Bt resistance monitoring. There
are two major limiting factors to be resolved yet. The first is the fact that
diverse mechanisms of resistance may exist in field populations (Zhang
et al., 2012a), challenging the accurate prediction of different or novel
resistance mechanisms evolved in the field. Selection for resistance in the
laboratory is a routine approach to predict likely mechanisms, but the resis-
tance may not be representative of resistance found in the field. F2 screens
might be applicable for isolating a collection of resistant strains representing
312 Yidong Wu
Table 6.2 Status of field-evolved resistance to Bt crops in nine species of major insect
pests (Tabashnik et al., 2013, 2014)
Pest Crop Toxin Country References
Incipient resistance (<1% resistant individuals)
H. armigera Cotton Cry1Ac Australia Downes and Mahon (2012b)
H. armigera Cotton Cry2Ab Australia Downes and Mahon (2012b)
H. punctigera Cotton Cry2Ab Australia Downes et al. (2010a)
Early warning (16% resistant individuals)
H. zea Cotton Cry2Ab United Ali and Luttrell (2007) and Tabashnik
States et al. (2009a, 2013)
Practical resistance (>50% resistant individuals and reduced efficacy reported)
B. fusca Corn Cry1Ab South Van Rensburg (2007) and Kruger
Africa et al. (2011)
D.v. virgifera Corn Cry3Bb United Gassmann et al. (2011, 2012)
States
H. zea Cotton Cry1Ac United Luttrell et al. (2004), Ali et al. (2006)
States and Tabashnik et al. (2008)
P. gossypiella Cotton Cry1Ac India Dhurua and Gujar (2011)
S. frugiperda Corn Cry1F United Storer et al. (2010, 2012)
States
3. RESISTANCE MECHANISMS
Mode of action of Bt Cry toxins has been intensively studied and fre-
quently reviewed (Bravo et al., 2011; Ferre and Van Rie, 2002; Griffitts and
Aroian, 2005; Ibrahim et al., 2010; Knowles, 1994; Pardo-Lopez et al.,
2013; Pigott and Ellar, 2007; Soberon et al., 2009; Vachon et al., 2012;
Chapter 2). Although many details of mode of action of Bt Cry toxins
are far from understood, the major steps (crystal solubilization, proteolytic
activation, receptor binding, membrane insertion and pore formation) are
generally agreed on (Knowles, 1994; Schnepf et al., 1998). Insects could
develop resistance to Cry toxins due to alteration at any step of the sequential
procession of intoxication, and as many as 10 potential Bt resistance mech-
anisms were proposed (Heckel, 1994). For brevity, major steps of mode of
action, potential and observed resistance mechanisms are represented in
Fig. 6.2. This review focuses on resistance mechanism involving alterations
in receptor binding and proteolytic activation of Cry toxins in lepidopteran
insects. Other mechanisms such as more efficient repair of damaged midgut
cells and elevated immune responses in Bt-resistant strains are not included.
Figure 6.2 Schematic diagram of the steps of Cry toxin action and possible resistance
mechanisms. Receptor mutations and/or proteolytic alterations have been identified in
a number of Bt-resistant lepidopterans, whereas resistance mechanisms involving sol-
ubilization failure or abnormal membrane insertion/pore formation have not been
reported yet.
Since identification of the first cry genes (Schnepf and Whiteley, 1981),
more than 700 cry genes divided into 72 groups have been reported
(Crickmore et al., 2014). Cry toxins are grouped in three families that are
not related in structure: Bin-like, Mtx-like and 3d-Cry (Pardo-Lopez
et al., 2013). Although the sequence identity is low between different groups
of 3d-Cry proteins, the three-dimensional structure is highly conserved.
The structure of 3d-Cry toxins is composed of three structural domains
(Li et al., 1991). Domain I consists of a bundle of seven -helices with a cen-
tral helix -5 surrounded by six amphipathic helices. Domain I is the most
conserved domain among Cry toxins and is responsible for pore formation in
the cell membrane. Domain II is composed of three antiparallel -sheets
316 Yidong Wu
Figure 6.3 Types of cadherin mutations associated with Cry1Ac resistance in three
cotton pests (Heliothis virescens, Pectinophora gossypiella and Helicoverpa armigera).
(1) Truncation: a premature stop codon is present at any sites on the extracellular
domain of cadherin. (2) Deletion in the extracellular domain: a stretch of amino acid res-
idues on the extracellular domain, present within or ahead of the toxin-binding region.
(3) Deletion in the intracellular domain: lack of 55 amino acid residues in the r15 allele of
HaCad, which confers non-recessive resistance to Cry1Ac (Zhang et al., 2012b). (4) Amino
acid substitution: the L1425R mutant of HvCad of H. virescens can decrease binding to
Cry1Ac and has the potential to confer resistance (Xie et al., 2005). The r10r14 alleles of
HaCad in H. armigera were suggested to confer Cry1Ac resistance by amino acid sub-
stitutions (Zhang et al., 2012a).
3.3.2 Aminopeptidase
Mutations or altered expression of APN has been suggested or confirmed to
confer high levels of resistance to Cry1 toxins in three lepidopterans:
S. exigua, T. ni and H. armigera (Herrero et al., 2005; Tiewsiri and Wang,
2011; Zhang et al., 2009).
Lack of APN1 expression was detected in a Cry1Ca-resistant strain of
S. exigua by using suppression subtractive hybridization. Northern blot anal-
ysis confirmed APN1 was not expressed in the resistant strain, whereas other
three APNs (APN2, APN3 and APN4) had no difference in expression
between the resistant and susceptible strains (Herrero et al., 2005). These
data suggest that the lack of APN1 expression plays a role in Cry1Ca resis-
tance in S. exigua, although a linkage between Cry1Ca resistance and the
lack of APN1 expression needs to be determined. Cry1Ac resistance in
the GLEN-Cry1Ac-BCS of T. ni was introgressed into a susceptible strain
322 Yidong Wu
3.3.4 ABCC2
In the YHD2 strain of H. virescens, a mutation in the cadherin gene HvCad
was identified as a major mechanism conferring high levels of resistance to
Cry1Ac toxin (Gahan et al., 2001). YHD2 lost binding with Cry1Aa, but
not with Cry1Ab and Cry1Ac (Lee et al., 1995). The YHD2 strain was fur-
ther selected with Cry1Ac to produce the YHD3 strain, which had much
higher resistance levels to Cry1Ac than YHD2 and loss binding with all three
Cry1A toxins ( Jurat-Fuentes and Adang, 2004). A second resistance gene
(HvABCC2) contributing to the additional resistance and lost of binding
with Cry1Ab and Cry1Ac in YHD3 was then identified by genetic mapping
and positional cloning (Gahan et al., 2010). A 22-bp deletion in exon 2 of
HvABCC2 in YHD3 was predicted to result in a truncated 99-residue pro-
tein instead of the full-length protein composed of 1339 amino acids. The
inactivation mutation of HvABCC2 was correlated with both higher
resistance levels and the loss of Cry1Ab and Cry1Ac binding to midgut
membranes (Gahan et al., 2010).
Resistance to Cry1Ac spray formulations has also evolved in field
populations of P. xylostella and T. ni ( Janmaat and Myers, 2003;
Tabashnik et al., 1990). Genetic mapping demonstrated that field-evolved
resistance to Cry1Ac in the NO-QA strain of P. xylostella and the
GLEN-Cry1Ac-BCS strain of T. ni was also genetically linked with ABCC2
gene, named PxABCC2 and TnABCC2, respectively (Baxter et al., 2011).
In the NO-QA strain of P. xylostella, a 10-residue deletion in the middle of
transmembrane helix 12 of PxABCC2 was predicted to be located in the
extracellular region of the second NBD domain (Baxter et al., 2011). How-
ever, the mutation of TnABCC2 has not yet been identified in the GLEN-
Cry1Ac-BCS strain of T. ni.
In the C2 strain of B. mori, a candidate locus for recessive resistance to
Cry1Ab was mapped to a region of 82 kb (containing BmABCC2) on chro-
mosome 15. Comparisons of BmABCC2 sequences among 10 susceptible
and 7 resistant strains revealed a common tyrosine insertion in resistant
alleles which is located in an outer loop of the predicted transmembrane
structure. Introduction of a wild-type allele of BmABCC2 into a resistant
strain by using germline transformation restored susceptibility to Cry1Ab
in the resistant strain, which provides very strong evidence for the role of
324 Yidong Wu
Figure 6.4 Diverse genetic basis of Cry1Ac resistance in field populations of Helicoverpa
armigera from China (Zhang et al., 2012a). Survival at the diagnostic concentration of
Cry1Ac of nine resistant (R) strains (red bars; eight strains isolated from the F2 screen
and the laboratory-selected strain SCD-r1 with the r1 allele of HaCad at the cadherin locus),
and progeny from crosses between each resistant strain and either the susceptible SCD
strain (green bars) or the resistant SCD-r1 strain (blue bars). Asterisks indicate 0% survival
for progeny from crosses between the SCD strains. Rec, recessive; Dom, dominant.
cadherin locus (Zhang et al., 2012a). In contrast with diverse genetic basis of
Cry1Ac resistance in H. armigera from China, five Cry2Ab-resistant strains of
H. armigera, which were isolated from an F2 screen of field populations in
Australia, were all recessive and shared a common resistance locus
(Mahon et al., 2008).
as resistance to more than one class of pesticides in a single pest through dif-
ferent resistance mechanisms. Multiple resistance is generally caused by
sequential or mosaic uses of different classes of pesticides. For example, a
strain of P. xylostella (NO-95C), selected in the field with Bt formulations
(containing Cry1A and Cry1C toxins) and further selected with Cry1C
in the laboratory, developed high levels of resistance to Cry1A and moderate
levels of resistance to Cry1C (Liu and Tabashnik, 1997; Liu et al., 1996).
Bioassays of progeny from split broods of single-pair families demonstrated
that resistance to Cry1C segregates independently of resistance to Cry1Ab in
NO-95C (Liu and Tabashnik, 1997). Resistance to Cry1A and Cry1C in
the NO-95C strain of P. xylostella is a typical phenomenon of multiple
resistance.
The pyramid strategy for Bt resistance management uses crops that
produces two or more dissimilar toxins that kill the same target pest
(Roush, 1998). Toxins to be used for pyramiding should have no cross-
resistance, so that the risk to evolve a single mechanism of resistance against
both toxins could be dramatically reduced. Mathematical models and empir-
ical studies have suggested resistance to pyramided two-gene plants that
show no cross-resistance can be significantly delayed as compared with resis-
tance to single-gene plants adopted sequentially or in mosaics (Roush, 1998;
Zhao et al., 2003).
However, target pests can possibly evolve a single gene that can over-
come both Bt genes to be used in the pyramid, even if they have different
binding sites. The most widely used pyramid is the second-generation Bt
cotton that produces Cry1Ac and Cry2Ab. No cross-resistance between
these two toxins is presumed because they bind to different larval midgut
target sites. However, several strains of H. virescens selected with Cry1Ac
in the laboratory caused significant cross-resistance to Cry2Aa (Gould
et al., 1992, 1995; Jurat-Fuentes et al., 2003). In addition, laboratory selec-
tion of a strain of P. gossypiella with Cry2Ab resulted in 420-fold cross-
resistance to Cry1Ac as well as 240-fold resistance to Cry2Ab (Tabashnik
et al., 2009b). Also, two Cry1Ac-resistant strains of H. armigera isolated from
field-selected populations from northern China had low, but significant
cross-resistance to Cry2Ab (46-fold; Jin et al., 2013), and a strain of
H. zea selected for resistance to Cry1Ac increased survival on Bt cotton
expressing Cry1Ac and Cry2Ab (Brevaulta et al., 2013). All these results
show that cross-resistance occurs between Cry1Ac and Cry2Ab in at least
three key cotton pests. Even though there are no shared binding sites
between two Cry proteins, there are several potential resistance mechanisms
Insect Resistance to Bt Cry Toxins 331
that could result in cross-resistance in the target insects. For example, if dif-
ferent classes of Bt toxins are activated by the same proteases, alterations of
the proteases could cause cross-resistance. If a novel proteolytic mechanism
is evolved to degrade the activated toxin core, it could cause cross-resistance.
Pore formation is believed to be general and an important step in the mech-
anism of action of many Cry proteins, thus any change affecting pore for-
mation or pore function may result in broad cross-resistance (Heckel,
1994). With increasing deployment of the pyramid Bt cotton and corn
crops, more work is needed to identify potential molecular basis of cross-
resistance between two distinct toxins such as Cry1Ac and Cry2Ab. In
the future, the toxins for stacking varieties should be as disparate as possible,
showing highly different mechanism of actions such as a Bt toxin and a non-
Bt factor, to minimize the chance that mutations in a single gene could con-
fer resistance to both factors.
5. CONCLUSIONS
Large-scale adoption of Bt crops have brought both ecological benefits
for environment and economic benefits for farmers. In the foreseeable
future, adoption of Bt crops will keep increasing globally. Benefited from
implementation of proactive resistance management strategies such as the
refuge strategy and the pyramid strategy in many countries, most of the tar-
get pests of Bt crops have been sustainably and effectively controlled for
nearly 20 years. However, several cases of field-evolved resistance to Bt corn
and Bt cotton have been documented causing reduced field efficacy. Evo-
lution of resistance by target pests is a real threat to the continued success of
Bt crops.
The genetic capacity of insect populations to evolve resistance to Cry
toxins has been well demonstrated by both laboratory-selected and field-
evolved resistance in many species within several insect orders. Diverse
genetic options for target pests to cope with Bt crops are challenging the
efforts to understand mechanisms of resistance and to design rational resis-
tance management strategies. Genetic mapping approach has been proved to
be a powerful tool and it is anticipated that it will continue to be essential in
dissecting the complex genetic basis of Bt resistance. Clarifying resistance
mechanisms can facilitate and advance our understanding on modes of action
of Cry toxins.
Proactive evaluation of the inheritance and initial frequency of resistance
are useful for assessing the risk of resistance problems in the close future.
332 Yidong Wu
ACKNOWLEDGEMENTS
I am grateful to Yihua Yang, Derek Russell, Alejandra Bravo, Mario Soberon, Bruce
Tabashnik and the editors of this volume for reading the text and making helpful
comments. This work was supported in part by grants from the Ministry of Agriculture of
China (Grant no. 2014ZX08012-004) and the NSFC of China (Grant no. 31272382).
REFERENCES
Alcantara, E., Estrada, A., Alpuerto, V., Head, G., 2011. Monitoring Cry1Ab susceptibility in
Asian corn borer (Lepidoptera: Crambidae) on Bt corn in the Philippines. Crop. Prot.
30, 554559.
Ali, M.I., Luttrell, R.G., 2007. Susceptibility of bollworm and tobacco budworm
(Noctuidae) to Cry2Ab2 insecticidal protein. J. Econ. Entomol. 100, 921931.
Ali, M.I., Luttrell, R.G., Young III, S.Y., 2006. Susceptibility of Helicoverpa zea and Heliothis
virescens (Lepidoptera: Noctuidae) populations to Cry1Ac insecticidal protein. J. Econ.
Entomol. 99, 164175.
Andow, D.A., Alstad, D.N., 1998. F2 screen for rare resistance alleles. J. Econ. Entomol.
91, 572578.
Andow, D.A., Alstad, D.N., Pang, Y.-H., Bolin, P.C., Hutchinson, W.D., 1998. Using an
F2 screen to search for resistance alleles to Bacillus thuringiensis toxin in European corn
borer (Lepidoptera: Crambidae). J. Econ. Entomol. 91, 579584.
Andow, D.A., Olson, D.M., Hellmich, R.L., Alstad, D.N., Hutchison, W.D., 2000. Fre-
quency of resistance to Bacillus thuringiensis toxin Cry1Ab in an Iowa population of
European corn borer (Lepidoptera: Crambidae). J. Econ. Entomol. 93, 2630.
Andreadis, S.S., Alvarez-Alfageme, F., Sanchez-Ramos, I., Stodola, T.J., Andow, D.A.,
Milonas, P.G., Savopoulou-Soultani, M., Castanera, P., 2007. Frequency of resistance
to Bacillus thuringiensis toxin CrylAb in Greek and Spanish population of Sesamia
nonagrioides (Lepidoptera: Noctuidae). J. Econ. Entomol. 100, 195201.
Anilkumar, K.J., Rodrigo-Simon, A., Ferre, J., Pusztai-Carey, M., Sivasupramaniam, S.,
Moar, W.J., 2008. Production and characterization of Bacillus thuringiensis Cry1Ac-
Insect Resistance to Bt Cry Toxins 333
Caprio, M.A., Sumerford, D.V., 2007. Evaluating transgenic plants for suitability in pest and
resistance management problems. In: Lacey, L.A., Kaya, H.K. (Eds.), Field Manual of
Techniques in Invertebrate Pathology. Springer, Netherlands, pp. 769789.
Carriere, Y., Crowder, D.W., Tabashnik, B.E., 2010. Evolutionary ecology of adaptation to
Bt crops. Evol. Appl. 3, 561573.
Carroll, J., Ellar, D.J., 1993. An analysis of Bacillus thuringiensis -endotoxin action on insect-
midgut-membrane permeability using a light-scattering assay. Eur. J. Biochem.
214, 771778.
Chen, J., Aimanova, K.G., Fernandez, L.E., Bravo, A., Soberon, M., Gill, S.S., 2009a. Aedes
aegypti cadherin serves as a putative receptor of the Cry11Aa toxin from Bacillus thur-
ingiensis subsp. israelensis. Biochem. J. 424, 191200.
Chen, J., Aimanova, K.G., Pan, S., Gill, S.S., 2009b. Identification and characterization of
Aedes aegypti aminopeptidase N as a putative receptor of Bacillus thuringiensis Cry11A
toxin. Insect Biochem. Mol. Biol. 39, 688696.
Contreras, E., Schoppmeier, M., Real, M.D., Rausell, C., 2013. Sodium solute symporter
and cadherin proteins act as Bacillus thuringiensis Cry3Ba toxin functional receptors in Tri-
bolium castaneum. J. Biol. Chem. 288, 1801318021.
Crava, C.M., Bel, Y., Jakubowska, A.K., Ferre, J., Escriche, B., 2013. Midgut aminopepti-
dase N isoforms from Ostrinia nubilalis: activity characterization and differential binding
to Cry1Ab and Cry1Fa proteins from Bacillus thuringiensis. Insect Biochem. Mol. Biol.
43, 924935.
Crickmore, N., Baum, J., Bravo, A., Lereclus, D., Narva, K., Sampson, K., Schnepf, E.,
Sun, M., Zeigler, D.R., 2014. Bacillus thuringiensis toxin nomenclature. http://www.
btnomenclature.info/.
Crow, J.F., 1960. Genetics of insecticide resistance: general considerations. Misc. Publ.
Entomol. Soc. Am. 2, 6974.
de Maagd, R.A., Bravo, A., Crickmore, N., 2001. How Bacillus thuringiensis has evolved spe-
cific toxins to colonize the insect world. Trends Genet. 17, 193199.
Devos, Y., Meihls, L.N., Kiss, J., Hibbard, B.E., 2013. Resistance evolution to the first gen-
eration of genetically modified Diabrotica-active Bt-maize events by western corn
rootworm: management and monitoring considerations. Transgenic Res. 22, 269299.
Dhurua, S., Gujar, G.T., 2011. Field-evolved resistance to Bt toxin Cry1Ac in the pink boll-
worm, Pectinophora gossypiella (Saunders) (Lepidoptera: Gelechiidae), from India. Pest
Manag. Sci. 67, 898903.
Downes, S., Mahon, R., 2012a. Evolution, ecology and management of resistance in Hel-
icoverpa spp. to Bt cotton in Australia. J. Invertebr. Pathol. 110, 281286.
Downes, S., Mahon, R., 2012b. Successes and challenges of managing resistance in Hel-
icoverpa armigera to Bt cotton in Australia. GM Crops Food 3, 228234.
Downes, S., Parker, T.L., Mahon, R.J., 2009. Frequency of alleles conferring resistance to the
Bacillus thuringiensis toxins Cry1Ac and Cry2Ab in Australian populations of Helicoverpa
punctigera (Lepidoptera: Noctuidae) from 2002 to 2006. J. Econ. Entomol. 102, 733742.
Downes, S., Parker, T., Mahon, R., 2010a. Incipient resistance of Helicoverpa punctigera to the
Cry2Ab Bt toxin in Bollgard II(r) cotton. PLoS One 5, e12567.
Downes, S., Parker, T.L., Mahon, R.J., 2010b. Characteristics of resistance to Bacillus thur-
ingiensis toxin Cry2Ab in a strain of Helicoverpa punctigera (Lepidoptera: Noctuidae) iso-
lated from a field population. J. Econ. Entomol. 103, 21472154.
Fabrick, J.A., Tabashnik, B.E., 2012. Similar genetic basis of resistance to Bt toxin Cry1Ac in
boll-selected and diet-selected strains of pink bollworm. PLoS One 7, e35658.
Fabrick, J., Oppert, C., Lorenzen, M.D., Morris, K., Oppert, B., Jurat-Fuentes, J.L., 2009.
A novel Tenebrio molitor cadherin is a functional receptor for Bacillus thuringiensis Cry3Aa
toxin. J. Biol. Chem. 284, 1840118410.
Insect Resistance to Bt Cry Toxins 335
Fabrick, J.A., Ponnuraj, J., Singh, A., Tanwar, R.K., Unnithan, G.C., Yelich, A.J., Li, X.C.,
Carriere, Y., Tabashnik, B.E., 2014. Alternative splicing and highly variable cadherin
transcripts associated with field-evolved resistance of pink bollworm to Bt cotton in
India. PLoS One 9, e97900.
Ferre, J., Van Rie, J., 2002. Biochemistry and genetics of insect resistance to Bacillus thur-
ingiensis. Annu. Rev. Entomol. 47, 501533.
Ferre, J., Van Rie, J., MacIntosh, S.C., 2008. Insecticidal genetically modified crops and
insect resistance management (IRM). In: Romeis, J., Shelton, A.M., Kennedy, G.G.
(Eds.), Integration of Insect-Resistant Genetically Modified Crops Within IPM Pro-
grams. Springer, Netherlands, pp. 4185.
ffrench-Constant, R.H., 2013. The molecular genetics of insecticide resistance. Genetics
194, 807815.
ffrench-Constant, R.H., Roush, R.T., 1990. Resistance detection and documentation: the
relative roles of pesticidal and biochemical assays. In: Roush, R.T., Tabashnik, B.E.
(Eds.), Pesticide Resistance in Arthropods. Chapman and Hall, New York, pp. 438.
Flannagan, R.D., Yu, C.G., Mathis, J.P., Meyer, T.E., Shi, X., Siqueira, H.A., Siegfried, B.D.,
2005. Identification, cloning and expression of a Cry1Ab cadherin receptor from European
corn borer, Ostrinia nubilalis (Hubner) (Lepidoptera: Crambidae). Insect Biochem. Mol.
Biol. 35, 3340.
Forrester, N.W., Cahill, M., Bird, L.J., Layland, J.K., 1993. Management of pyrethroid and
endosulfan resistance in Helicoverpa armigera (Lepidoptera: Noctuidae) in Australia. Bull.
Entomol. Res. Suppl. Ser. 1, 1132.
Gahan, L.J., Gould, F., Heckel, D.G., 2001. Identification of a gene associated with Bt resis-
tance in Heliothis virescens. Science 293, 857860.
Gahan, L.J., Gould, F., Lopez Jr., J.D., Micinski, S., Heckel, D.G., 2007. A polymerase chain
reaction screen of field populations of Heliothis virescens for a retrotransposon insertion
conferring resistance to Bacillus thuringiensis toxin. J. Econ. Entomol. 100, 187194.
Gahan, L.J., Pauchet, Y., Vogel, H., Heckel, D.G., 2010. An ABC transporter mutation is
correlated with insect resistance to Bacillus thuringiensis Cry1Ac toxin. PLoS Genet.
6, e1001248.
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.
Gassmann, A.J., Petzold-Maxwell, J.L., Keweshan, R.S., Dunbar, M.W., 2012. Western
corn rootworm and Bt maize: challenges of pest resistance in the field. GM Crops Food
3, 235244.
Genissel, A., Augustin, S., Courtin, C., Pilate, G., Lorme, P., Bourguet, D., 2003. Initial
frequency of alleles conferring resistance to Bacillus thuringiensis poplar in a field popula-
tion of Chrysomela tremulae. Proc. Biol. Sci. 270, 791797.
Gill, S.S., Cowles, E.A., Francis, V., 1995. Identification, isolation, and cloning of a Bacillus
thuringiensis CryIAc toxin-binding protein from the midgut of the lepidopteran insect
Heliothis virescens. J. Biol. Chem. 270, 2727727282.
Gill, M., Ellar, D., 2002. Transgenic Drosophila reveals a functional in vivo receptor for the
Bacillus thuringiensis toxin Cry1Ac1. Insect Mol. Biol. 11, 619625.
Gomez, I., Sanchez, J., Munoz-Garay, C., Matus, V., Gill, S.S., Soberon, M., Bravo, A.,
2014. Bacillus thuringiensis Cry1A toxins are versatile proteins with multiple modes of
action: two distinct pre-pores are involved in toxicity. Biochem. J. 459, 383396.
Gould, F., 1998. Sustainability of transgenic insecticidal cultivars: integrating pest genetics
and ecology. Annu. Rev. Entomol. 43, 701726.
Gould, F., Martinez-Ramirez, A., Anderson, A., Ferre, J., Silva, F.J., Moar, W.J., 1992.
Broad-spectrum resistance to Bacillus thuringiensis toxins in Heliothis virescens. Proc. Natl.
Acad. Sci. U.S.A. 89, 79867990.
336 Yidong Wu
Gould, F., Anderson, A., Reynolds, A., Bumgarner, L., Moar, W.J., 1995. Selection and
genetic analysis of a Heliothis virescens (Lepidoptera: Noctuidae) strain with high levels
of resistance to Bacillus thuringiensis toxins. J. Econ. Entomol. 88, 15451559.
Gould, F., Anderson, A., Jones, A., Sumerford, D., Heckel, D.G., Lopez, J., Micinski, S.,
Leonard, R., Laster, M., 1997. Initial frequency of alleles for resistance to Bacillus thur-
ingiensis toxins in field populations of Heliothis virescens. Proc. Natl. Acad. Sci. U.S.A.
94, 35193523.
Griffitts, J.S., Aroian, R.V., 2005. Many roads to resistance: how invertebrates adapt to Bt
toxins. Bioessays 27, 614624.
Groeters, F.R., Tabashnik, B.E., 2000. Roles of selection intensity, major genes, and minor
genes in evolution of insecticide resistance. J. Econ. Entomol. 93, 15801587.
Heckel, D.G., 1994. The complex genetic basis of resistance to Bacillus thuringiensis toxin in
insects. Biocontrol Sci. Technol. 4, 405417.
Heckel, D.G., 2012. Learning the ABCs of Bt: ABC transporters and insect resistance to
Bacillus thuringiensis provide clues to a crucial step in toxin mode of action. Pestic. Bio-
chem. Physiol. 104, 103110.
Heckel, D.G., Gahan, L.C., Gould, F., Anderson, A., 1997. Identification of a linkage group
with a major effect on resistance to Bacillus thuringiensis Cry1Ac endotoxin in the tobacco
budworm (Lepidoptera: Noctuidae). J. Econ. Entomol. 90, 7586.
Heckel, D.G., Gahan, L.J., Baxter, S.W., Zhao, J.Z., Shelton, A.M., Gould, F.,
Tabashnik, B.E., 2007. The diversity of Bt resistance genes in species of Lepidoptera.
J. Invertebr. Pathol. 95, 192197.
Herrero, S., Oppert, B., Ferre, J., 2001. Different mechanisms of resistance to Bacillus thur-
ingiensis toxins in the indianmeal moth. Appl. Environ. Microbiol. 67, 10851089.
Herrero, S., Gechev, T., Bakker, P.L., Moar, W.J., de Maagd, R.A., 2005. Bacillus thur-
ingiensis Cry1Ca-resistant Spodoptera exigua lacks expression of one of four aminopepti-
dase N genes. BMC Genomics 6, 96.
Hua, G., Zhang, R., Abdullah, M.A., Adang, M.J., 2008. Anopheles gambiae cadherin
AgCad1 binds the Cry4Ba toxin of Bacillus thuringiensis israelensis and a fragment of
AgCad1 synergizes toxicity. Biochemistry 47, 51015110.
Hua, G., Zhang, R., Bayyareddy, K., Adang, M.J., 2009. Anopheles gambiae alkaline phospha-
tase is a functional receptor of Bacillus thuringiensis jegathesan Cry11Ba toxin.
Biochemistry 48, 97859793.
Hua, G., Zhang, Q., Zhang, R., Abdullah, A.M., Linser, P.J., Adang, M.J., 2013. AgCad2
cadherin in Anopheles gambiae larvae is a putative receptor of Cry11Ba toxin of Bacillus
thuringiensis subsp. jegathesan. Insect Biochem. Mol. Biol. 43, 153161.
Huang, F., 2006. Detection and monitoring of insect resistance to transgenic Bt crops. Insect
Sci. 13, 7384.
Huang, F., Leonard, B.R., Cook, D.R., Lee, D.R., Andow, D.A., Baldwin, J.L., Tindall, K.V.,
Wu, X., 2007a. Frequency of alleles conferring resistance to Bacillus thuringiensis maize in
Louisiana populations of the southwestern corn borer. Entomol. Exp. Appl. 122, 5358.
Huang, F.N., Leonard, B.R., Andow, D.A., 2007b. F2 screen for resistance to a Bacillus thur-
ingiensis-maize hybrid in the sugarcane borer (Lepidoptera: Crambidae). Bull. Entomol.
Res. 97, 437444.
Huang, F.N., Leonard, B.R., Moore, S.H., Cook, D.R., Baldwin, J., Tindall, K.V., Lee, D.R.,
2008. Allele frequency of resistance to Bacillus thuringiensis Cry1Ab corn in Louisiana
populations of sugarcane borer (Lepidoptera: Crambidae). J. Econ. Entomol. 101, 492498.
Huang, F., Parker, R., Leonard, R., Yong, Y., Liu, J., 2009. Frequency of resistance alleles to
Bacillus thuringiensis-corn in Texas populations of the sugarcane borer, Diatraea saccharalis
(F.) (Lepidoptera: Crambidae). Crop. Prot. 28, 174180.
Huang, F., Andow, D.A., Buschman, L.L., 2011. Success of the high-dose/refuge resistance
management strategy after 15 years of Bt crop use in North America. Entomol. Exp.
Appl. 140, 116.
Insect Resistance to Bt Cry Toxins 337
Huang, F., Ghimire, M.N., Leonard, B.R., Daves, C., Levy, R., Baldwin, J., 2012. Extended
monitoring of resistance to Bacillus thuringiensis Cry1Ab maize in Diatraea saccharalis
(Lepidoptera: Crambidae). GM Crops Food 3, 245254.
Ibrahim, M.A., Griko, N., Junker, M., Bulla, L.A., 2010. Bacillus thuringiensis: a genomics and
proteomics perspective. Bioeng. Bugs 1, 3150.
James, C., 2013. Global Status of Commercialized Biotech/GMCrops: 2013: ISAAA Brief
No. 46. International Service for the Acquisition of Agri-biotech Applications (ISAAA),
Ithaca, NY.
Janmaat, A.F., Myers, J., 2003. Rapid evolution and the cost of resistance to Bacillus thur-
ingiensis in greenhouse populations of cabbage loopers, Trichoplusia ni. Proc. Biol. Sci.
270, 22632270.
Jimenez, A.I., Reyes, E.Z., Cancino-Rodezno, A., Bedoya-Perez, L.P., Caballero-Flores,
G.G., Muriel-Millan, L., Likitvivatanavong, F.S., Gill, S.S., Bravo, A., Soberon, M.,
2012. Aedes aegypti alkaline phosphatase ALP1 is a functional receptor of Bacillus thur-
ingiensis Cry4Ba and Cry11Aa toxins. Insect Biochem. Mol. Biol. 42, 683689.
Jin, L., Yin, W., Zhang, L., Yang, Y., Tabashnik, B.E., Wu, Y., 2013. Dominant resistance
to Bt cotton and minor cross-resistance to Bt toxin Cry2Ab in cotton bollworm from
China. Evol. Appl. 6, 12221235.
Jurat-Fuentes, J.L., Adang, M.J., 2004. Characterization of a Cry1Ac-receptor alkaline phos-
phatase in susceptible and resistant Heliothis virescens larvae. Eur. J. Biochem.
271, 31273135.
Jurat-Fuentes, J.L., Adang, M.J., 2006. The Heliothis virescens cadherin protein expressed in
Drosophila S2 cells functions as a receptor for Bacillus thuringiensis Cry1A but not Cry1Fa
toxins. Biochemistry 45, 96889695.
Jurat-Fuentes, J.L., Gould, F.L., Adang, M.J., 2003. Dual resistance to Bacillus thuringiensis
Cry1Ac and Cry2Aa toxins in Heliothis virescens suggests multiple mechanisms of resis-
tance. Appl. Environ. Microbiol. 69, 58985906.
Jurat-Fuentes, J.L., Karumbaiah, L., Jakka, S.R.K., Ning, C.M., Liu, C.X., Wu, K.M.,
Jackson, J., Gould, F., Blanco, C., Portilla, M., Perera, O., Adang, M., 2011. Reduced
levels of membrane-bound alkaline phosphatase are common to lepidopteran strains
resistant to Cry toxins from Bacillus thuringiensis. PLoS One 6, e17606.
Khakame, S.K., Wang, X., Wu, Y., 2013. Baseline toxicity of metaflumizone and lack of
cross resistance between indoxacarb and metaflumizone in diamondback moth
(Lepidoptera: Plutellidae). J. Econ. Entomol. 106, 14231429.
Knight, P.J., Crickmore, N., Ellar, D.J., 1994. The receptor for Bacillus thuringiensis CrylA(c)
delta-endotoxin in the brush border membrane of the lepidopteran Manduca sexta is ami-
nopeptidase N. Mol. Microbiol. 11, 429436.
Knight, P.J., Knowles, B.H., Ellar, D.J., 1995. Molecular cloning of an insect
aminopeptidase-N that serves as a receptor for Bacillus thuringiensis CryIA(c) toxin.
J. Biol. Chem. 270, 1776517770.
Knowles, B.H., 1994. Mechanism of action of Bacillus thuringiensis insecticidal delta-
endotoxins. Adv. Insect Physiol. 24, 275308.
Knowles, B.H., Ellar, D.J., 1987. Colloid-osmotic lysis is a general feature of the mechanism
of action of Bacillus thuringiensis -endotoxins with different insect specificity. Biochim.
Biophys. Acta 924, 507518.
Kruger, M., Van Rensburg, J.B.J., Van Den Berg, J., 2011. Resistance to Bt maize in Busseola
fusca (Lepidoptera: Noctuidae) from Vaalharts, South Africa. Environ. Entomol.
40, 477483.
Lee, M.K., Rajamohan, F., Gould, F., Dean, D.H., 1995. Resistance to Bacillus thuringiensis
Cry1A delta-endotoxins in a laboratory-selected Heliothis virescens strain is related to
receptor alteration. Appl. Environ. Microbiol. 61, 38363842.
Li, J., Carrol, J., Ellar, D.J., 1991. Crystal structure of insecticidal delta-endotoxin from Bacil-
lus thuringiensis at 2.5 A resolution. Nature 353, 815821.
338 Yidong Wu
Liang, G.M., Wu, K.M., Yu, H.K., Li, K.K., Feng, X., Guo, Y.Y., 2008. Changes of inher-
itance mode and fitness in Helicoverpa armigera (Hubner) (Lepidoptera: Noctuidae) along
with its resistance evolution to Cry1Ac toxin. J. Invertebr. Pathol. 97, 142149.
Liu, Y.B., Tabashnik, B.E., 1997. Inheritance of resistance to the Bacillus thuringiensis toxin
Cry1C in the diamondback moth. Appl. Environ. Microbiol. 63, 22182223.
Liu, Y.B., Tabashnik, B.E., Pusztai-Carey, M., 1996. Field-evolved resistance to Bacillus thur-
ingiensis toxin Cry1C in diamondback moth (Lepidoptera: Plutellidae). J. Econ.
Entomol. 89, 798804.
Liu, F., Xu, Z., Zhu, Y.C., Huang, F., Wang, Y., Li, H.L., Li, H., 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, 155161.
Luo, K., Sangadala, S., Masson, L., Mazza, A., Brousseau, R., Adang, M.J., 1997. The
Heliothis virescens 170-kDa aminopeptidase functions as Receptor A by mediating
specific Bacillus thuringiensis Cry1A -endotoxin binding and pore formation. Insect Bio-
chem. Mol. Biol. 27, 735743.
Luttrell, R.G., Ali, I., Allen, K.C., Young III, S.Y., Szalanski, A., Williams, K., Lorenz, G.,
Parker Jr., C.D., Blanco, C., 2004. Resistance to Bt in Arkansas populations of cotton
bollworm. In: Richter, D.A. (Ed.), Proceedings, 2004 Beltwide Cotton Conferences,
59 January 2004, San Antonio, TX. National Cotton Council of America,
Memphis, TN, pp. 13731383.
Mahon, R.J., Olsen, K.M., Downes, S., Addison, S., 2007. Frequency of alleles conferring
resistance to the Bt toxins Cry1Ac and Cry2Ab in Australian populations of Helicoverpa
armigera. J. Econ. Entomol. 100, 18441853.
Mahon, R.J., Olsen, K.M., Downes, S., 2008. Isolations of Cry2Ab resistance in Australian
populations of Helicoverpa armigera (Lepidoptera: Noctuidae) are allelic. J. Econ. Entomol.
101, 909914.
Mahon, R.J., Downes, S., James, W., Parker, T., 2010. Why do F1 screens estimate higher
frequencies of Cry2Ab resistance in Helicoverpa armigera (Lepidoptera: Noctuidae) than do
F2 screens? J. Econ. Entomol. 103, 472481.
Mahon, R.J., Downes, S.J., James, B., 2012. Vip3A resistance alleles exist at high levels in
Australian targets before release of cotton expressing this toxin. PLoS One 7, e39192.
Martins, E.S., Monnerat, R.G., Queiroz, P.R., Dumas, V.F., Braz, S.V., de Souza Aguiar,
R.W., Gomes, A.C., Sanchez, J., Bravo, A., Ribeiro, B.M., 2010. Midgut GPI-
anchored proteins with alkaline phosphatase activity from the cotton boll weevil
(Anthonomus grandis) are putative receptors for the Cry1B protein of Bacillus thuringiensis.
Insect Biochem. Mol. Biol. 40, 138145.
McKenzie, J.A., Batterham, P., 1994. The genetic, molecular and phenotypic consequences
of selection for insecticide resistance. Trends Ecol. Evol. 9, 166169.
McNall, R.J., Adang, M.J., 2003. Identification of novel Bacillus thuringiensis Cry1Ac binding
proteins in Manduca sexta midgut through proteomic analysis. Insect Biochem. Mol. Biol.
33, 9991010.
Moar, W.J., Pusztai-Carey, M., Van Faassen, H., Bosch, D., Frutos, R., Rang, C., Luo, K.,
Adang, M.J., 1995. Development of Bacillus thuringiensis CryIC resistance by Spodoptera
exigua (Hubner) (Lepidoptera: Noctuidae). Appl. Environ. Microbiol. 61, 20862092.
Moar, W., Roush, R., Shelton, A., Ferre, J., MacIntosh, S., Leonard, B.R., Abel, C., 2008.
Field-evolved resistance to Bt toxins. Nat. Biotechnol. 26, 10721074.
Morin, S., Biggs, R.W., Sisterson, M.S., Shriver, L., Ellers-Kirk, C., Higginson, D.,
Holley, D., Gahan, L.J., Heckel, D.G., Carriere, Y., 2003. Three cadherin alleles asso-
ciated with resistance to Bacillus thuringiensis in pink bollworm. Proc. Natl. Acad. Sci.
U.S.A. 100, 50045009.
Nagamatsu, Y., Toda, S., Koike, T., Miyoshi, Y., Shigematsu, S., Kogure, M., 1998. Clon-
ing, sequencing, and expression of the Bombyx mori receptor for Bacillus thuringiensis
insecticidal CryIA(a) toxin. Biosci. Biotechnol. Biochem. 62, 727734.
Insect Resistance to Bt Cry Toxins 339
Ning, C.M., Wu, K.M., Liu, C.X., Gao, Y.L., Jurat-Fuentes, J.L., Gao, J.L., 2010. Char-
acterization of a Cry1Ac toxin-binding alkaline phosphatase in the midgut from Hel-
icoverpa armigera (Hubner) larvae. J. Insect Physiol. 56, 666672.
Oltean, D.I., Pullikuth, A.K., Lee, H.-K., Gill, S.S., 1999. Partial purification and
characterization of Bacillus thuringiensis Cry1A toxin receptor A from Heliothis
virescens and cloning of the corresponding cDNA Appl. Environ. Microbiol.
65, 47604766.
Oppert, B., 1999. Protease interactions with Bacillus thuringiensis insecticidal toxins. Arch.
Insect Biochem. Physiol. 42, 112.
Oppert, B., Kramer, K.J., Beeman, R.W., Johnson, D., McGaughey, W.H., 1997.
Proteinase-mediated insect resistance to Bacillus thuringiensis toxins. J. Biol. Chem.
272, 2347323476.
Pardo-Lopez, L., Soberon, M., Bravo, A., 2013. Bacillus thuringiensis insecticidal three-
domain Cry toxins: mode of action, insect resistance and consequences for crop protec-
tion. FEMS Microbiol. Rev. 37, 322.
Perera, O.P., Willis, J.D., Adang, M.J., Jurat-Fuentes, J.L., 2009. Cloning and characteriza-
tion of the Cry1Ac-binding alkaline phosphatase (HvALP) from Heliothis virescens. Insect
Biochem. Mol. Biol. 39, 294302.
Perlak, F.J., Deaton, R.W., Armstrong, T.A., Fuchs, R.L., Sims, S.R., Greenplate, J.T.,
Fischhoff, D.A., 1990. Insect resistant cotton plants. Nat. Biotechnol. 8, 939943.
Petzold-Maxwell, J.L., Cibils-Stewart, X., Wade French, B., Gassmann, A.J., 2012. Adap-
tation by western corn rootworm (Coleoptera: Chrysomelidae) to Bt maize: inheritance,
fitness costs, and feeding preference. J. Econ. Entomol. 105, 14071418.
Pigott, C.R., Ellar, D.J., 2007. Role of receptors in Bacillus thuringiensis crystal toxin activity.
Microbiol. Mol. Biol. Rev. 71, 255281.
Rajagopal, R., Sivakumar, S., Agrawal, N., Malhotra, P., Bhatnagar, R.K., 2002. Silencing
of midgut aminopeptidase N of Spodoptera litura by double stranded RNA establishes its
role as Bacillus thuringiensis toxin receptor. J. Biol. Chem. 277, 4684946851.
Rajagopal, R., Arora, N., Sivakumar, S., Rao, N.G., Nimbalkar, S.A., Bhatnagar, R.K.,
2009. Resistance of Helicoverpa armigera to Cry1Ac toxin from Bacillus thuringiensis is
due to improper processing of the protoxin. Biochem. J. 419, 309316.
Ren, X.L., Chen, R.R., Zhang, Y., Ma, Y., Cui, J.J., Han, Z.J., Mu, L.L., Li, G.Q., 2013.
A Spodoptera exigua cadherin serves as a putative receptor for Bacillus thuringiensis Cry1Ca
toxin and shows differential enhancement of Cry1Ca and Cry1Ac toxicity. Appl. Envi-
ron. Microbiol. 79, 55765583.
Roush, R.T., 1998. Two-toxin strategies for management of insecticidal transgenic crops:
can pyramiding succeed where pesticide mixtures have not? Philos. Trans. R. Soc. B
353, 17771786.
Roush, R.T., McKenzie, J.A., 1987. Ecological genetics of insecticide and acaricide resis-
tance. Annu. Rev. Entomol. 32, 361380.
Roush, R.T., Miller, G.L., 1986. Considerations for design of insecticide resistance moni-
toring. J. Econ. Entomol. 79, 293298.
Sanahuja, G., Banakar, R., Twyman, R.M., Capell, T., Christou, P., 2011. Bacillus thur-
ingiensis: a century of research, development and commercial applications. Plant Bio-
technol. J. 9, 283300.
Sawicki, R., 1987. Definition, detection and documentation of insecticide resistance.
In: Ford, M.G., Holloman, D.W., Khambay, B.P.S., Sawicki, R.M. (Eds.), Combating
Resistance to Xenobiotics: Biological and Chemical Approaches. Ellis Horwood Ltd.,
Chichester, UK, pp. 105117
Sayed, A., Nekl, E.R., Siqueira, H.A., Wang, H.C., ffrench-Constant, R.H., Bagley, M.,
Siegfried, B.D., 2007. A novel cadherin-like gene from western corn rootworm,
Diabrotica virgifera virgifera (Coleoptera: Chrysomelidae), larval midgut tissue. Insect
Mol. Biol. 16, 591600.
340 Yidong Wu
Schnepf, H.E., Whiteley, H.R., 1981. Cloning and expression of the Bacillus thuringiensis crystal
protein gene in Escherichia coli. Proc. Natl. Acad. Sci. U.S.A. 78, 28932897.
Schnepf, E., Crickmore, N., Van Rie, J., Lereclus, D., Feitelson, J., Zeigler, D.R., Dean, D.H.,
1998. Bacillus thuringiensis and its pesticidal crystal proteins. Microbiol. Mol. Biol. Rev.
62, 775806.
Schwartz, J.L., Potvin, L., Chen, X.J., Brousseau, R., Laprade, R., Dean, D.H., 1997.
Single-site mutations in the conserved alternating arginine region affect ionic channels
formed by CryIAa, a Bacillus thuringiensis toxin. Appl. Environ. Microbiol.
63, 39783984.
Siegfried, B.D., Rangasamy, M., Wang, H., Spencer, T., Haridas, C.V., Tenhumberg, B.,
Sumerford, D.V., Storer, N.P., 2014. Estimating the frequency of Cry1F resistance in
field populations of the European corn borer (Lepidoptera: Crambidae). Pest Manag.
Sci. 70, 725733.
Sivakumar, S., Rajagopal, R., Venkatesh, G.R., Srivastava, A., Bhatnagar, R.K., 2007.
Knockdown of aminopeptidase-N from Helicoverpa armigera larvae and in transfected
Sf21 cells by RNA interference reveals its functional interaction with Bacillus thuringiensis
insecticidal protein Cry1Ac. J. Biol. Chem. 282, 73127319.
Sivasupramaniam, S., Head, G.P., English, L., Li, Y.J., Vaughn, T.T., 2007. A global
approach to resistance monitoring. J. Invertebr. Pathol. 95, 224226.
Soberon, M., Gill, S.S., Bravo, A., 2009. Signaling versus punching hole: how do Bacillus
thuringiensis toxins kill insect midgut cells? Cell. Mol. Life Sci. 66, 13371349.
Stodola, T.J., Andow, D.A., Hyden, A.R., Hinton, J.L., Roark, J.J., Buschman, L.L.,
Porter, P., Cronholm, G.B., 2006. Frequency of resistance to Bacillus thuringiensis toxin
Cry1Ab in southern United States corn belt population of European corn borer
(Lepidoptera: Crambidae). J. Econ. Entomol. 99, 502507.
Storer, N.P., Babcock, J.M., Schlenz, M., Meade, T., Thompson, G.D., Bing, J.W.,
Huckaba, R.M., 2010. Discovery and characterization of field resistance to Bt maize:
Spodoptera frugiperda (Lepidoptera: Noctuidae) in Puerto Rico. J. Econ. Entomol.
103, 10311038.
Storer, N.P., Kubiszak, M.E., King, J.E., Thompson, G.D., Santos, A.C., 2012. Status of
resistance to Bt maize in Spodoptera frugiperda: lessons from Puerto Rico. J. Invertebr. Pat-
hol. 110, 294300.
Sumerford, D.V., Head, G.P., Shelton, A., Greenplate, J., Moar, W., 2013. Field-evolved
resistance: assessing the problem and moving forward. J. Econ. Entomol. 106, 15251534.
Tabashnik, B.E., 1994. Evolution of resistance to Bacillus thuringiensis. Annu. Rev. Entomol.
39, 4779.
Tabashnik, B.E., Cushing, N.L., Finson, N., Johson, M.W., 1990. Field development of
resistance to Bacillus thuringiensis in diamondback moth (Lepidoptera: Plutellidae).
J. Econ. Entomol. 83, 16711676.
Tabashnik, B.E., Finson, N., Johnson, M.W., Moar, W.J., 1993. Resistance to toxins from
Bacillus thuringiensis subsp. kurstaki causes minimal cross-resistance to B. thuringiensis
subsp. aizawai in the diamondback moth (Lepidoptera: Plutellidae). Appl. Environ.
Microbiol. 59, 13321335.
Tabashnik, B.E., Liu, Y.B., Dennehy, T.J., Sims, M.A., Sisterson, M.S., Biggs, R.W.,
Carriere, Y., 2002. Inheritance of resistance to Bt toxin Cry1Ac in a field-derived strain
of pink bollworm (Lepidoptera: Gelechiidae). J. Econ. Entomol. 95, 10181026.
Tabashnik, B.E., Gould, F., Carriere, Y., 2004. Delaying evolution of insect resistance to
transgenic crops by decreasing dominance and heritability. J. Evol. Biol. 17, 904912.
Tabashnik, B.E., Fabrick, J.A., Henderson, S., Biggs, R.W., Yafuso, C.M., Nyboer, M.E.,
Manhardt, N.M., Coughlin, L.A., Sollome, J., Carriere, Y., Dennehy, T.J., Morin, S.,
2006. DNA screening reveals pink bollworm resistance to Bt cotton remains rare after a
decade of exposure. J. Econ. Entomol. 99, 15251530.
Insect Resistance to Bt Cry Toxins 341
Tabashnik, B.E., Gassmann, A.J., Crowder, D.W., Carriere, Y., 2008. Insect resistance to Bt
crops: evidence versus theory. Nat. Biotechnol. 26, 199202.
Tabashnik, B.E., Van Rensburg, J.B.J., Carriere, Y., 2009a. Field-evolved insect resistance to
Bt crops: definition, theory, and data. J. Econ. Entomol. 102, 20112025.
Tabashnik, B.E., Unnithan, G.C., Masson, L., Crowder, D.W., Li, X., Carriere, Y., 2009b.
Asymmetrical cross-resistance between Bacillus thuringiensis toxins Cry1Ac and Cry2Ab
in pink bollworm. Proc. Natl. Acad. Sci. U.S.A. 106, 1188911894.
Tabashnik, B.E., Sisterson, M.S., Ellsworth, P.C., Dennehy, T.J., Antilla, L., Liesner, L.,
Whitlow, M., Staten, R.T., Fabrick, J.A., Unnithan, G.C., Yelich, A.J., Ellers-Kirk,
C., Harpold, V.S., Li, X., Carriere, Y., 2010. Suppressing resistance to Bt cotton with
sterile insect releases. Nat. Biotechnol. 28, 13041307.
Tabashnik, B.E., Huang, F.N., Ghimire, M.N., Leonard, B.R., Siegfried, B.D.,
Rangasamy, M., Yang, Y.J., Wu, Y.D., Gahan, L.J., Heckel, D.G., Bravo, A.,
Soberon, M., 2011. Efficacy of genetically modified Bt toxins against insects with differ-
ent genetic mechanisms of resistance. Nat. Biotechnol. 29, 11281131.
Tabashnik, B.E., Brevault, T., Carriere, Y., 2013. Insect resistance to Bt crops: lessons from
the first billion acres. Nat. Biotechnol. 31, 510521.
Tabashnik, B.E., Mota-Sanchez, D., Whalon, M.E., Hollingworth, R.M., Carriere, Y.,
2014. Defining terms for proactive management of resistance to Bt crops and pesticides.
J. Econ. Entomol. 107, 496507.
Tanaka, S., Miyamoto, K., Noda, H., Jurat-Fuentes, J.L., Yoshizawa, Y., Endo, H., Sato, R.,
2013. The ATP-binding cassette transporter subfamily C member 2 in Bombyx mori lar-
vae is a functional receptor for Cry toxins from Bacillus thuringiensis. FEBS J.
280, 17821794.
Tiewsiri, K., Wang, P., 2011. Differential alteration of two aminopeptidases N associated
with resistance to Bacillus thuringiensis toxin Cry1Ac in cabbage looper. Proc. Natl. Acad.
Sci. U.S.A. 108, 1403714042.
Vachon, V., Laprade, R., Schwartz, J.L., 2012. Current models of the mode of action of Bacil-
lus thuringiensis insecticidal crystal proteins: a critical review. J. Invertebr. Pathol.
111, 112.
Vadlamudi, R.K., Weber, E., Ji, I., Ji, T.H., Bulla Jr., L.A., 1995. Cloning and expression of
a receptor for an insecticidal toxin of Bacillus thuringiensis. J. Biol. Chem.
270, 54905494.
Van Rensburg, J.B.J., 2007. First report of field resistance by stem borer, Busseola fusca (Fuller)
to Bt-transgenic maize. S. Afr. J. Plant Soil 24, 147151.
Wan, P., Huang, Y., Wu, H., Huang, M., Cong, S., Tabashnik, B.E., Wu, K., 2012.
Increased frequency of pink bollworm resistance to Bt toxin Cry1Ac in China. PLoS
One 7, e29975.
Wang, X., Li, X., Shen, A., Wu, Y., 2010. Baseline susceptibility of the diamondback moth
(Lepidoptera: Plutellidae) to chlorantraniliprole in China. J. Econ. Entomol.
103, 843848.
Wenes, A.L., Bourguet, D., Andow, D.A., Courtin, C., Carre, G., Lorme, P., Sanchez, L.,
Augustin, S., 2006. Frequency and fitness cost of resistance to Bacillus thuringiensis in
Chrysomela tremulae (Coleoptera: Chrysomelidae). Heredity 97, 127134.
WHO, 1957. Expert Committee on Malaria, Seventh Report: WHO Technical Report
Series No. 125. World Health Organization, Geneva, Switzerland.
Xie, R., Zhuang, M., Ross, L.S., Gomez, I., Oltean, D.I., Bravo, A., Soberon, M., Gill, S.S.,
2005. Single amino acid mutations in the cadherin receptor from Heliothis virescens affect
its toxin binding ability to Cry1A toxins. J. Biol. Chem. 280, 84168425.
Xu, X., Yu, L., Wu, Y., 2005. Disruption of a cadherin gene associated with resistance to
Cry1Ac -endotoxin of Bacillus thuringiensis in Helicoverpa armigera. Appl. Environ.
Microbiol. 71, 948954.
342 Yidong Wu
Xu, Z., Liu, F., Chen, J., Huang, F., Andow, D.A., Wang, Y., Zhu, Y.C., Shen, J., 2009.
Using an F2 screen to monitor frequency of resistance alleles to Bt cotton in field
populations of Helicoverpa armigera (Hubner) (Lepidoptera: Noctuidae). Pest Manag.
Sci. 65, 391397.
Yang, Y.J., Chen, H., Wu, S., Xu, X., Wu, Y., 2006. Identification and molecular detection
of a deletion mutation responsible for a truncated cadherin of Helicoverpa armigera. Insect
Biochem. Mol. Biol. 36, 735740.
Yang, Y.J., Chen, H., Wu, Y., Yang, Y.-H., Wu, S., 2007. Mutated cadherin alleles from a
field population of Helicoverpa armigera confer resistance to Bacillus thuringiensis toxin
Cry1Ac. Appl. Environ. Microbiol. 73, 69396944.
Yang, Y.H., Yang, Y.J., Gao, W.Y., Guo, J.J., Wu, Y.H., Wu, Y.D., 2009. Introgression of a
disrupted cadherin gene enables susceptible Helicoverpa armigera to obtain resistance to
Bacillus thuringiensis toxin Cry1Ac. Bull. Entomol. Res. 99, 175181.
Yang, Y., Zhu, Y.C., Ottea, J., Husseneder, C., Leonard, B.R., Abel, C., Luttrell, R.,
Huang, F., 2011. Down regulation of a gene for cadherin, but not alkaline phosphatase,
associated with Cry1Ab resistance in the sugarcane borer Diatraea saccharalis. PLoS One
6, e25783.
Yue, B., Huang, F., Leonard, B.R., Moore, S., Parker, R., Andow, D.A., Cook, D.,
Emfinger, K., Lee, D.R., 2008. Verifying an F1 screen for identification and quantifica-
tion of rare Bacillus thuringiensis resistance alleles in field populations of the sugarcane
borer, Diatraea saccharalis. Entomol. Exp. Appl. 129, 172180.
Zhang, X., Candas, M., Griko, N.B., Rose-Young, L., Bulla, L.A., 2005. Cytotoxicity of
Bacillus thuringiensis Cry1Ab toxin depends on specific binding of the toxin to the
cadherin receptor BT-R1 expressed in insect cells. Cell Death Differ. 12, 14071416.
Zhang, X., Candas, M., Griko, N.B., Taussig, R., Bulla, L.A., 2006. A mechanism of cell
death involving an adenylyl cyclase/PKA signaling pathway is induced by the Cry1Ab
toxin of Bacillus thuringiensis. Proc. Natl. Acad. Sci. U.S.A. 103, 98979902.
Zhang, S., Cheng, H., Gao, Y., Wang, G., Liang, G., Wu, K., 2009. Mutation of an ami-
nopeptidase N gene is associated with Helicoverpa armigera resistance to Bacillus thur-
ingiensis Cry1Ac toxin. Insect Biochem. Mol. Biol. 39, 421429.
Zhang, H., Yin, W., Zhao, J., Jin, L., Yang, Y., Wu, S., Tabashnik, B.E., Wu, Y., 2011.
Early warning of cotton bollworm resistance associated with intensive planting of Bt cot-
ton in China. PLoS One 6, e22874.
Zhang, H., Tian, W., Zhao, J., Jin, L., Yang, J., Liu, C., Yang, Y., Wu, S., Wu, K., Cui, J.,
Tabashnik, B.E., Wu, Y., 2012a. Diverse genetic basis of field-evolved resistance to Bt
cotton in cotton bollworm from China. Proc. Natl. Acad. Sci. U.S.A. 109, 1027510280.
Zhang, H., Wu, S., Yang, Y., Tabashnik, B.E., Wu, Y., 2012b. Non-recessive Bt toxin resis-
tance conferred by an intracellular cadherin mutation in field-selected populations of cot-
ton bollworm. PLoS One 7, e53418.
Zhang, H., Tang, M., Yang, F., Yang, Y., Wu, Y., 2013. DNA-based screening for an intra-
cellular cadherin mutation conferring non-recessive Cry1Ac resistance in field
populations of Helicoverpa armigera. Pestic. Biochem. Physiol. 107, 148152.
Zhao, J.Z., Cao, J., Li, Y., Collins, H.L., Roush, R.T., Earle, E.D., Shelton, A.M., 2003.
Transgenic plants expressing two Bacillus thuringiensis toxins delay insect resistance evo-
lution. Nat. Biotechnol. 21, 14931497.
Zhao, J., Jin, L., Yang, Y., Wu, Y., 2010. Diverse cadherin mutations conferring resistance to
Bacillus thuringiensis toxin Cry1Ac in Helicoverpa armigera. Insect Biochem. Mol. Biol.
40, 113118.
Zuniga-Navarrete, F., Gomez, I., Pena, G., Bravo, A., Soberon, M., 2013. A Tenebrio molitor
GPI-anchored alkaline phosphatase is involved in binding of Bacillus thuringiensis Cry3Aa
to brush border membrane vesicles. Peptides 41, 8186.
CHAPTER SEVEN
Photorhabdus Toxins
Richard H. ffrench-Constant, Andrea J. Dowling
Biosciences, University of Exeter, Cornwall, United Kingdom
Contents
1. Photorhabdus Lifestyles, Relatives and Genomes 344
1.1 The life cycle of Photorhabdus temperata and Photorhabdus luminescens 344
1.2 The unusual life cycle of Photorhabdus asymbiotica 345
1.3 Xenorhabdus and comparative genomics 346
2. The Toxin Complexes 347
2.1 Tc discovery, gene cloning and ABC nomenclature 348
2.2 Diversity of tc-like genes from other bacteria 349
2.3 Structure, function and biophysics of the Tc ABC complexes 350
2.4 The Tcs as polymorphic toxins 361
2.5 The role of the Tcs in the biology of infection 361
2.6 The potential role of Tcs in crop protection 363
3. Photorhabdus Virulence Cassettes 364
3.1 Discovery and organization of PVCs 365
3.2 The role of PVC-like structures in other bacteria 367
3.3 Implications for PVC biology 369
4. The Mcf Toxins 370
4.1 Discovery and mode of action of Mcf1 370
4.2 Diversity of Mcf-like toxins 372
4.3 Studying Mcf-like toxins in vivo 374
5. Patox and Photox 376
5.1 PaTox structure and function 376
5.2 Photox as a novel mART toxin 378
6. Binary Toxins 379
6.1 The PirAB binary toxins 379
6.2 The XaxAB and YaxAB cytotoxins 379
7. Classical Secretions Systems and Novel Screens 380
7.1 Type III and other classical secretion systems 380
7.2 RVA-like screens for novel effectors 381
7.3 Clustering methods to look for novel effector phenotypes 382
8. Perspectives for the Future of Photorhabdus Toxins 383
Acknowledgements 383
References 383
Abstract
The last 10 years has seen an explosion in our knowledge of the diversity of toxins pro-
duced by the insect pathogenic bacteria Photorhabdus and Xenorhabdus. Here we
review new data on the structure, mode of action and biophysics of Photorhabdus
toxins and place them into context with similar toxins made in other bacteria such
as those of the genus Yersinia and Serratia. We look in detail at structurefunction stud-
ies of the Toxin complexes or Tcs and suggest that these might constitute a new mode
of bacterial secretion, here termed the Toxin complex secretion system or TC-SS. We
examine current data on the Makes Caterpillars Floppy or Mcf toxins both from insect
pathogenic bacteria and also the related Mcf-like Fit toxin from Pseudomonas. We also
review data on the Photorhabdus Virulence Cassettes or PVCs and their equivalent
phage-like secretion systems in other bacteria. Again these appear to be a totally novel
secretion system, termed PVC-SS, derived from tailed myophages. Finally, we review the
sheer diversity of other candidate toxins identified from Photorhabdus and Xenorhabdus
genomes, such as PaTox and Photox, and from simple gain-of-function screens devised
to identify novel effectors when expressed in recombinant Escherichia coli, termed
Rapid Virulence Annotation or RVA. Taken together, available data support the early
contention that Photorhabdus genomes do indeed encode more candidate toxins than
those of other bacteria and suggest that this unique genus demands more time from
serious toxinologists than has been devoted to date.
the insects hemocoel does not make it obvious as to why any Photorhabdus
toxins are indeed active on the insect midgut, unless they are somehow able
to exert their effects from either side of the gut (gut lumen or insect hemo-
coel). This section, therefore, describes the discovery, purification, cloning
and structure of the toxins conferring this oral activity, namely the Toxin
complexes or Tcs, and finally attempts reconcile the mode of action of these
important toxins with their potential role in infection.
and tcaC (encoding B and C) in the same bacterial cytoplasm was also essen-
tial (Waterfield et al., 2001a), whereas the tcaA open reading frame
(encoding A) could be expressed alone and then mixed with the products
of tcaB/tcaC (B and C) to reconstitute full oral toxicity (Waterfield et al.,
2001a). To put it most simply, a mixture of A, B and C toxin subunits
was required and, for some mysterious reason, B and C needed to be
made together in the same bacterial cytoplasm. This mechanistic explana-
tion of the A, B and C subunits also provided an instant cure to the growing
nomenclature problem as the massive list of gene names derived for different
tc loci within the same genome or the genomes of different bacteria, such
as Xenorhabdus, Serratia entomophila, Yersinia entomophaga and Yersinia pestis,
could all be resolved down to whether they encoded an A, B or
C subunit of these ABC toxins.
fact in a different order due to the split nature of the first gene in the
operon. Interestingly, and importantly, in some species of bacteria genes
encoding B and C subunits are found as gene fusions. In other words,
both the B and C subunits are encoded together in the same polypeptide.
In fact, Dow AgroSciences still holds patents that protect the fusion of
B and C encoding genes for their expression in transgenic plants (see
Section 2.6 on the use of Tc genes in crop protection below). This clearly
suggests that the B and C subunits must have some very close association in
formation of the final ABC complex, a point to which we will return below.
et al., 2001b). The central core of a tccC gene encodes a highly conserved
motif with similarity to sequences previously termed rearrangement hotspots or
rhs elements in E. coli. The N-terminal sequence is also well conserved but
the C-terminal tail is extremely variable. This variability of the C-termini of
tccC-like genes is found across all tccC genes in one genome or indeed across
all tccC-like genes in any bacterial genome. Critically, the laboratory of Klaus
Aktories has recently shown that two of these C subunits (termed TccC3
and TccC5) are ADP-ribosyltransferases that selectively modify unusual
amino acids with actin and Rho proteins, respectively (Gatsogiannis
et al., 2013; Lang et al., 2010, 2011). Specifically, TccC3 ADP-ribosylates
threonine-148 in actin, resulting in actin polymerization, and TccC5 ADP-
ribosylates Rho at glutamine-61/63, inducing Rho activation (Gatsogiannis
et al., 2013), and the net effect of these two enzymatic activities is to cause
extensive polymerization and clustering of actin in target cells. Intriguingly,
given that the active site of the ribosyltransferase is encoded within the
C-termini of the C subunits confirms the hypothesis that the C-termini
of the C subunits encode a range of different potential effectors
(Waterfield et al., 2001b). The tccC genes that encode the C subunits are,
therefore, a potential gold mine for the identification of further enzymatic
activities and translocated effectors. For example, the C-terminal end of one
C subunit from Y. pestis (termed YipA) carries a consensus sequence for pro-
tein tyrosine phosphatase (Spinner et al., 2012), suggesting one possible fur-
ther mode of enzymatic action. All of these data suggest a model whereby
the A subunits of the Tc toxins come together to form a syringe-like injec-
tion mechanism that somehow translocates the C-terminus of the C subunit
into the target cell where it exerts its toxic effect. This hypothesis is consis-
tent with data showing rearrangement of the A subunits upon membrane
insertion, suggesting that they do indeed act like a syringe for insertion
of the toxic moiety into the host cell. However, the precise interaction
of the B and C subunits and how it leads to the release of the C-terminus
of the C subunit awaited further detailed structural studies, this time again
from the Yersinia toxin.
protecting that host bacterium from any of the toxic effects that the
C subunit might exert on the host itself. The shell of the egg has a
-propeller domain at one end which it uses to attach to the A subunit in
the mature native ABC complex. Moreover, when the C subunit is folded
in complex with B, it auto-proteolyses to release its toxic C-terminus for
translocation into the target cell (Busby et al., 2013). This correlates
with the earlier observation that the Photorhabdus C subunit, specifically
TccC3, is cleaved at its hyper-variable sequence but only in the presence
of the B subunit, in this case TcdB2. This superb piece of work in
Y. entomophaga finally explains why the central core of the tccC-like genes
is so conserved, as it encodes the precise site at which auto-proteolysis occurs
and it must therefore always remain invariant for the C subunit cleavage to
work. The structure of the C subunit is also important as it is the first struc-
ture determined for proteins containing Rearrangement hotspot (Rhs)
repeats, similar structures to which are found in the eukaryotic tyrosine
aspartate YD-repeat-containing family, including the teneurins which are
type II integral membrane proteins that help establish neuronal cell connec-
tions during development (Hong et al., 2012). This may suggest that the Tc
B/C-mediated encapsulation device is used in other systems for protein
encapsulation and delivery. Indeed, recent work has suggested that
Gram-negative Rhs proteins and the distantly related wall-associated protein
A (WapA) from Gram-positive bacteria mediate intercellular competition in
bacteria. Specifically, Rhs and WapA carry polymorphic C-terminal toxin
domains that are deployed to inhibit the growth of neighbouring bacterial
cells (Kung et al., 2012). Similarly, RhsT from Pseudomonas aeruginosa PSE9
is translocated into mouse phagocytic cells and mice survive infection with
rhsT mutants, showing that RhsT is a real virulence factor (Kung et al.,
2012). These striking findings show that YD-repeat proteins in bacteria
are involved in contact-dependent growth inhibition and/or virulence,
and suggests that B/C-like YD-repeat containing egg shell structures could
play a more general role in protecting host bacterial cells from their delivered
bacterial toxins.
Figure 7.1 Ribbon diagram showing the TcA protomer in (A) pre-pore state and (B) in
pore-forming state. Receptor regions AD and neuramidase-like region are labelled,
along with the Tcb-binding domain, linker, pore-forming domain and -helical domains
(the small and large lobes). Note the dramatic changes that occur as the TcA protomer
switches from pre-pore to pore-forming states, specifically, the collapse of the elastic
linker, the extension of the pore-forming domain and the re-organization of the recep-
tor domains AD. Reprinted from Meusch et al. (2014) with permission from MacMillan.
Figure 7.2 Another view of the structure of (A) Tca pre-pore and (B) pore-forming com-
plex. Note the entropic spring or linker (black), release of which provides the elastic
energy to drive the pore-forming channel down towards the membrane. The
neuramidase-like domain is shown in blue (see text for discussion). Reprinted from
Meusch et al. (2014) with permission from MacMillan.
(Meusch et al., 2014). They propose that this hostile environment might be
able to directly unfold native proteins and suggest that the lack of resolution
of the cleaved ADP-ribosyltransferase domain residing in the shell is consis-
tent with it being either unfolded or in static disorder. This leaves us with a
picture of the unfolded ADP-ribosyltransferase domain sitting within the
B/C shell and awaiting translocation into the delivery channel formed by
the A subunits. So how does it get into the channel? The -propeller of
the B subunit interacts strongly with the highly conserved central funnel
formed by the A subunits, as shown in Fig. 7.3AC. The narrow gate
formed by this -propeller is open in the holotoxin, suggesting that it forms
the entrance to the delivery channel, as depicted in Fig. 7.3B. The open
-propeller pore is also hydrophobic and the authors suggest it therefore
plays a similar role to the -clamp in anthrax toxin which protects hydro-
phobic patches in the translocated protein.
The narrow passage of the -propeller suggests that the ADP-
ribosyltransferase must be unfolded before it passes this gate and critically
the authors found extra density (labelled in red in Fig. 7.4B) within the trans-
location channel after holotoxin formation, confirming the idea that the
unfolded ADP-ribosyltransferase passes through the -propeller gate and
enters the translocation channel before membrane permeation. Finally,
before we leave this description of a stunning new delivery mechanism, it
is worth returning to the unexplained diversity of potential toxins encoded
by the C-termini of different TccC-like proteins. Meusch et al. (2014) point
out that the C-terminal (hyper-variable) regions of TccC3 and TccC5 are
both cationic with isoelectric points of 9.68 and 8.65, respectively
(Meusch et al., 2014). This suggests that translocation of the ADP-
ribosyltransferase domains of TccC3 and TccC5, and indeed equivalent
domains from other unstudied TccC-like proteins are pH-independent. This
is in stark contrast to the pH gradient-dependent, unidirectional, transloca-
tion of anthrax lethal factor across the cation-selective anthrax translocation
pore (Brown et al., 2011). This also suggests that the other cleaved toxic
C-termini of other C subunits should also be cationic (have similar isoelectric
points) to take advantage of the same delivery mechanism.
Figure 7.3 Interaction of the B/C (TcB/TcC) subunits with the A (TcA) subunit assembly.
(A) The B subunit (labelled TcB in blue) and the C subunit (TcC in yellow) interact to form
a large cocoon- or egg-shell-shaped structure. The B subunit also comprises the
A subunit-binding domain (labelled Tca-binding domain in light blue) which folds into
an asymmetric six-bladed -propeller. (B) A vertical section through the egg-shell-like
structure formed by B/C to show the upper and lower chambers and the pre-chamber
formed by the -propeller. Note that the gate out of the B/C chamber is closed. Subunit
domains are coloured: Tca-binding domain (light blue), TcB (dark blue), TcC (yellow) and
Tcb N-terminal region (red). (C) Cryo-EM structure of the ABC holotoxin. Subunits and
domains are labelled: TcA (grey) and TcB/TcC (yellow). The inset labelled TcBTcC shows
that the -propeller gate is open here. (D) Final molecular model of the ABC holotoxin
formed by fitting the crystal structure of the A subunits and the crystal structure of B/C
into the overall cryo-EM structure. Reprinted from Meusch et al. (2014) with permission
from MacMillan.
Photorhabdus Toxins 359
Figure 7.4 Vertical sections through the density of the ABC holotoxin channel.
(A) Section showing the continuous channel between the egg-shell-like structure of
B/C in yellow and the channel of the A subunits in grey. (B) Difference map between
the cryo-EM structure of the ABC holotoxin and the cryo-EM structure of the
A subunits alone. The difference map is overlaid in red. Note the extra density present
in the channel which may correspond to the delivered ADP-ribosyltransferase domain
of the C subunit (see text above for discussion). Reprinted from Meusch et al. (2014) with
permission from MacMillan.
How are such large proteins extruded through such a small pore and exactly
how are they unfolded and re-folded? How are the mature holotoxins
released from the bacterial outer membrane to be found free in the bacterial
supernatant? In relation to the first question, it is interesting that the ABC
complexes of Yersinia appear to be decorated with chitinases, whereas no
similar companions have been noted for Tc holotoxins from Photorhabdus
or Xenorhabdus. There are clearly several potential and simple explanations
for the lack of these co-inhabitants in Photorhabdus and Xenorhabdus. (1) It
may simply be that the operons studied in Photorhabdus and Xenorhabdus lack
nearby genes encoding chitinases. There is, therefore, the formal possibility
that chitinases will form partners in ABC complexes if encoded within, or
nearby, the tc-encoding locus itself. (2) The potential work of the chitinases
as chaperones (i.e. the presence of chitinases is required for correct folding
and assembly of the mature ABC toxin) might be performed by other part-
ners in other bacteria. To this end, it is noteworthy that we recovered
GroEL-related amino acid sequences from our original purification of
native P. luminescens W14 Toxin complex A (D. Bowen, M. Blackburn
360 Richard H. ffrench-Constant and Andrea J. Dowling
natural history of most bacteria is far from complete and we are therefore only
left to speculate on the role of tc-like genes based on the bacteria for which
some insect-associated natural history is available. For example, in the well-
studied insect pathogen S. entomophila, the pADAP plasmid carrying A,
B and C subunit-encoding tc-like genes (sepA, sepB and sepC) clearly does
indeed cause amber disease in grass grubs (Tan et al., 2006). Similarly, three
tc-like genes are also found in Y. entomophaga, suggesting that the resultant
ABC toxin could have effects on the C. zealandica midgut (Hurst et al.,
2011). In fact, examination of the midgut histopathology of C. zealandica
treated with purified Yersinia Tc toxins (Marshall et al., 2012) is strikingly sim-
ilar to the effects seen on feeding purified native Photorhabdus holotoxin A to
M. sexta caterpillars (Blackburn et al., 1998). Specifically, first the peritrophic
membrane that surrounds the food in the gut is destroyed and then the micro-
villi of the midgut itself are disrupted. While such studies make for elegant
cross sections of the insect midgut, they may be failing to tell us much about
how the Tc toxins are really attaching to and destroying the midgut cells.
Thus, it is also noteworthy that the midgut histopathology of both Bt toxins
and indeed cholesterol oxidase (Purcell et al., 1993) are little different from
that of the Tc toxins. This histopathology may therefore simply reflect a com-
mon pattern of midgut self-destruction, whatever the proximal toxin or toxic
insult. However, taken together, the available data suggest that different insect
pathogens do indeed use the Tcs to disrupt the insect midgut, presumably
either from the gut lumen itself (Y. entomophaga or B. thuringiensis) or from
the hemocoel side (Photorhabdus and Xenorhabdus). Indeed, some pathogens
that carry tc-like genes, such as Y. entomophaga, have been shown to be able
to pass from the gut lumen into the hemocoel itself (Marshall et al., 2012),
suggesting the tc-like genes may somehow facilitate this passage across the gut.
Insects can also act as reservoirs or vectors of human diseases. To this end,
the role of tc-like genes in the plague bacillus Y. pestis has, therefore, attracted
considerable recent attention. In this case, Y. pestis is normally maintained by
flearodent enzootic cycles but is occasionally transferred to man through
the bite of an infected flea. Originally when the first genomes of Y. pestis
were sequenced, these tc-like genes were found disrupted in some strains
leading to the original hypothesis that their loss may somehow be involved
in increased pathogenicity to man. More recently, the genes encoding the
A and B subunits (termed yitAB and yitC) and two C subunits (the tccC
homologues yipA and yipB) have been shown to be largely intact in the
genomes of most other sequenced Y. pestis strains, suggesting that they
may in fact play an undescribed role in interacting with either the flea vector
Photorhabdus Toxins 363
or incidental human host (Spinner et al., 2012). The currently available data
on exactly what role the Tc toxins might perform in the unusual lifecycle of
the plague bacillus are however potentially conflicting. If we ignore the sug-
gestion that Y. pestis Tcs are type III secreted (Gendlina et al., 2007), which
now seems hard to believe given that they effectively form their own secre-
tion machinery, and concentrate on data available for tc gene expression, we
can see several interesting points. First, Y. pestis tc genes (yitAB and yitC) are
up-regulated in J774A.1 macrophages. Second, repression of tc genes via
knock-out of their yitR regulator allowed for increased phagocytosis of
Y. pestis isolated directly from fleas (Vadyvaloo et al., 2010). Suggesting that
the tc genes may be playing some role in inhibiting phagocytosis of the flea-
borne bacteria by the macrophages. Third expression of yitR, the regulator
of Y. pestis tc expression, is up-regulated in the flea itself (Vadyvaloo et al.,
2010) but the Tc proteins themselves do not play a detectable role in flea
infection or the ability to produce a transmissible infection (Spinner et al.,
2012). Fourth, the Y. pestis Tc proteins (A subunit protein, YitA and
C subunit protein, YipA) themselves are highly expressed in the flea but
not at the same temperature (22 C) in the laboratory (Spinner et al.,
2012). Fifth, Tc (YitA and YipA) production is higher at lower temperature
(<22 C) and is cut-off at 37 C, the temperature the bacterium would
reach after being released into a human, but the proteins persist for over
9 h (Spinner et al., 2012). Finally, Y. pestis Tc proteins are localized to
the outer membrane of the bacterium (Spinner et al., 2012), like the Tcs
of Photorhabdus. Taking all of this available evidence together, this suggests
that Tc proteins are made within infected fleas and then displayed on the
outer membrane of Y. pestis as it is introduced into mammalian host upon
flea feeding. This role would be consistent with them resisting phagocytosis
via rodent or human phagocytes after their release from the flea. Such a role
is consistent with them playing a role in pathogenicity against mammalian
hosts but potentially via the nature of the delivered C subunit toxin (i.e.
in preventing phagocytosis) rather than via the loss of insect-specific tc loci,
as originally suggested by the sequencing of the first Y. pestis genome.
Unfortunately our detailed knowledge of the structure and function of the Tcs
has not come hand in hand with their successful expression in transgenic crops,
despite their obvious potential utility as insecticidal proteins. Here we will not
provide a detailed breakdown of the patent literature on this area, which is
considerable, but it is worth considering why the expression of these genes
in plants is so technically demanding. First, before it was understood that
all three A, B and C subunits needed to be expressed in plants, considerable
effort was placed into making plants that only expressed A subunits, with the
underlying assumption that they would kill insects. For example, as a model
system the A subunit from P. luminescens W14 was expressed in transgenic Ara-
bidopsis and its toxicity to neonate M. sexta was examined (Liu et al., 2003).
We now clearly know that there are a number of reasons as to why we would
not expect this experiment to work. First, M. sexta (Tobacco or Tomato
hornworm) does not feed on this plant by choice, so while Arabidopsis is a use-
ful model for testing the success of expressing a large gene (the intact tcdA
mRNA expressed was 8 kb) and resulting protein (TcdA at 283 kDa) in
plants, we would not expect the feeding results to be very meaningful. Sec-
ond, and most importantly, apart from the potential for pore formation, it is
hard to see why expression of the A subunits alone would be toxic in the
absence of B and C subunits, specifically the delivered C toxin, unless binding
and pore formation by A subunits alone are indeed enough to trigger oral tox-
icity. However, despite all this, bioassays of 234 T0 lines transformed with
construct pDAB7026 (osm-tcdA) identified nine transformed lines that caused
100% mortality in neonate M. sexta infested on transgenic Arabidopsis (Liu
et al., 2003). These data are therefore consistent with the hypothesis that bind-
ing and potential pore formation by A subunits alone are enough to trigger
insect mortality, at least in this model insect. However, full recombinant tox-
icity clearly requires full assembly of the ABC holotoxin which to our knowl-
edge has not been successfully achieved in a transgenic plant. The simplest
reason for this is that expression of all three relatively large genes encoding
A, B and C subunits (or expression of A with a B/C fusion) is still beyond
the limits of current plant transformation technology or that the plant is simply
not able to perform all of assembly/chaperone functions required to assemble a
functional Tc holotoxin.
vehicle from Photorhabdus that we will discuss, termed the phage-like PVCs.
Belonging to a group of phage-tail-like particles termed tailocins, PVCs are
a fascinating structure with a potentially very interesting biology.
Cnf ). The open reading frames encoding these putative effectors are also
regularly flanked by transposons, suggesting that the different effector genes
may be swapped between different PVC loci. This array of putative toxins
encoded at the end of each PVC locus, and the potential to swap these toxin-
encoding genes around, suggests that the PVCs also belong to the polymor-
phic toxin system super-group (Zhang et al., 2012), as discussed above for
the Tc-encoding genes. Zhang and co-workers have noted that several toxin
domains are secreted by the PVC-SS across most major bacterial lineages and
even the euryarchaea. Moreover, proteins secreted via the PVC-SS carry a
highly conserved metallopeptidase domain (identified by the motif
HExxHxxQ-E) N-terminal to the toxin domain itself. So again we see a
system, like the TccCs, whereby a constant N-terminal peptidase domain
can recombine to gather different variable C-terminal toxin domains (Zhang
et al., 2012) and again this suggests that these peptidase domains are likely to
act as auto-proteolytic domains that release the toxin during or after its secre-
tion by the PVC-SS. The C-terminal toxins found to be associated with
these PVC metalloproteases span an incredible diversity of nucleases, nucleic
acid deaminases, peptidases, pore-forming domains and several other enzy-
matic domains (Zhang et al., 2012).
PVCs can be readily expressed in recombinant E. coli and expression of
the proteins corresponding to the first three open reading frames of the
P. asymbiotica PVC carrying Photorhabdus necrosis factor (PaPVCpnf ) was
confirmed via 2D gel analysis and protein sequencing. Transmission electron
microscopy (TEM) of the recombinant PVCs revealed a phage-tail-like par-
ticle very similar to R-type pyocins (Yang et al., 2006). The PVCs are
30 nm wide but show a huge variety in length (up to 800 nm long)
depending on whether they are found in a contracted (needle within sheath)
or extended state (needle protruding from sheath). Interestingly, polyclonal
antibodies raised against the putative effector proteins do seem to label pro-
teins extruded from the PVC when anti-effector antibody-conjugated gold
particles are viewed under TEM (N.R. Waterfield and R.H. ffrench-
Constant, unpublished data). The candidate effector proteins also cause
extensive rearrangement of the actin cytoskeleton when expressed directly
inside transfected NIH-3T3 cells, consistent with them being bona fide effec-
tors. Finally, injection of larvae of the wax moth, Galleria mellonella, with
recombinant PaPVCpnf killed the larvae and caused dramatic rearrangement
of the actin cytoskeleton of hemocytes recovered from injected larvae (Yang
et al., 2006). Such effects disappeared when transposons were inserted into
the open reading frames encoding the putative effector proteins (Yang et al.,
Photorhabdus Toxins 367
T4 and also some components of the type VI secretion system. Using cryo-
EM and electron tomography of negatively stained Afp particles, Heymann
et al. (2013) have recently shown that the Afp sheath is similar to that of the
T4 phage tail but with a different sub-domain arrangement of the polymer-
izing sheath proteins. The central channel, which presumably delivers the
putative effectors encoded at the end of the PVC-like locus, is similar in
diameter and axial width to the Hcp1 hexamer of the P. aeruginosa type
VI secretion system. This tube extends through the baseplate into
a needle, resembling the puncture device of the T4 phage tail. Interest-
ingly, the tube itself appears to contain either the delivered toxin or poten-
tially a ruler protein which may measure the length of the tube while
under assembly. Taken together, the available Afp data support a hypothesis
whereby one of two different candidate effectors, also encoded on pADAP,
is delivered by the modified phage tail into target cells in the insect midgut.
At the time of writing, however, the exact receptors for the Afp particle,
how it binds to the midgut cell and also exactly how the toxin is delivered
all remain a mystery.
More recently, another stunning piece of bacterial natural history has
been described around a second PVC-like structure, the Mac, of the marine
bacterium P. luteoviolacea. Larvae of the serpulid polychaete tubeworm,
Hydroides elegans, require contact with surface-bound P. luteoviolacea bacteria
in order to undergo metamorphosis. This has recently been shown to be
determined by an array of PVC-like phage-tail-like structures that the bac-
terium makes. These Mac arrays are formed from 100 contractile struc-
tures with outward-facing baseplates, linked by tail fibres and a dynamic
hexagonal net (Shikuma et al., 2014). These arrays are synthesized intracel-
lularly, are released by cell lysis and then expand extracellularly into an
ordered multi-Mac array. Previously, Huang and co-authors had specu-
lated that the Macs might be involved in transporting specific molecules
either into the biofilm matrix or into larval cells directly that then leads
to larval metamorphosis (Huang et al., 2012). The presentation of the Macs
in an ordered array may therefore suggest that their firing or delivery is syn-
chronized or that the array somehow guarantees the correct dosage of deliv-
ered effector to trigger metamorphosis when contacted by a settling worm
larva. Either way, this incredible piece of biology supports the idea that a
PVC-like structure can deliver an unknown bacterial effector into a marine
worm larva that is then triggered to undergo metamorphosis. This piece of
biology reminds us that the relationship between bacteria, phage and
eukaryotes is very ancient and that bacterial effectors that we view simply
Photorhabdus Toxins 369
as toxins due to their effects on isolated cells may in fact be effectors that
modulate exquisite aspects of bacteriumanimal interactions. Thus in the
case of the PVCs expressed and released by Photorhabdus in the infected
insect, these phage-like delivery vehicles might even be interacting with
the infective (dauer) juvenile nematode hosts and triggering their develop-
ment (metamorphosis) into reproductive hermaphrodites.
How do PVC-like secretion systems recognize their target cells? To
begin to come up with some hypotheses to answer this question, its notable
that some tailed phage also carry virion proteins that show peptidoglycan
hydrolytic activity (Moak and Molineux, 2004), suggesting that these are
used by the phage to penetrate the peptidoglycan cell wall during the deliv-
ery of DNA. Phage P2 gpX also carries LysM domains which have been
implicated in binding to both peptidoglycan and chitin, the latter of which
would clearly be important in the recognition of insect tissue. In conclusion,
and at this point in time, we are clearly not certain as to how the released
Photorhabdus PVCs bind to or indeed penetrate host cell walls but, given
the above, it is interesting to speculate that they may be redeploying mech-
anisms evolved to get into prokaryotic cells (Gram-positive and Gram-
negative bacteria) in their interactions with eukaryotic cells.
had the ability to kill caterpillars and also caused apoptotic cell death in both
isolated insect hemocytes and also the midgut of injected caterpillars,
suggesting that Mcf may be promoting apoptosis (Daborn et al., 2002).
The mcf gene predicts a 2929-amino acid protein with a molecular
weight of 324 kDa. Homology searches suggested that this large protein
was divided into three overall domains (Daborn et al., 2002). First, the
N-terminal domain carries a domain similar to a Bcl2-homology 3-like
(BH3-like) domain (labelled as BH3 in Fig. 7.5), whose putative function
we will return to. Second, the central domain is similar to Clostridium difficile
toxin B over a region important for toxin translocation (labelled as Cyto-
toxin B-like in Fig. 7.5). Third, the C-terminus carries a region similar to
repeats in a Repeats in toxins (RTX)-like toxin from Actinobacillus
pleuropneumoniae (labelled as RTX in Fig. 7.5).
While none of these homologies are particularly striking, we wanted to
test the possibility that Mcf was promoting apoptosis via its similarity to a
BH3-domain-only protein. BH3-domain-only proteins are important in
the homeostatic balance of Bcl2 proteins in the outer membrane of the mito-
chondrion, and overexpression of BH3-domain-only proteins in mammals
causes mitochondrially mediated apoptosis (Budd, 2001). To test this
hypothesis, we looked at the effects of recombinant Mcf toxin on mamma-
lian cell lines. We found that Mcf-treated cells showed apoptotic nuclear
Figure 7.5 Diagram showing the inter-relationship between the different domains in
Mcf-like toxins. The Photorhabdus toxin Mcf1 carries similarities to the predicted
MCF1/SHE proteolysis domain (domain labelled SHE), a BH3-like domain (BH3) and
two regions that carry similarity to Repeats in toxins (RTX) toxins. The central domain
is similar to Clostridium difficile toxin B over a region important for toxin translocation
(labelled as Cytotoxin B-like). The Photorhabdus toxin Mcf2 has a different N-terminal
domain which carries similarity to the type III delivered plant effector HrmA (domain
labelled HrmA-like). Finally, the Mcf-like FitD toxin carries the SHE, BH3 and one of
the RTX-like domains but lacks similarity to RTX-like toxins at its C-terminus (see text
for further discussion).
372 Richard H. ffrench-Constant and Andrea J. Dowling
all of the injected larvae in 88 h. They also found that when they changed the
DxD motif in PaTox to asparagine-X-asparagine (NxN), 61% of the larvae
survived ( Jank et al., 2013). The authors also tested the role of the putative
glycosyltransferase domain (termed PaToxG) in inhibiting phagocytosis by
macrophages using protective antigen (the component of anthrax toxin
responsible for binding and translocation) as a delivery system. Strikingly,
PaToxG strongly inhibited phagocytosis of labelled E. coli, an effect that
was eliminated when the DxD motif was again changed to NxN. Similarly,
the PaToxG domain also caused disassembly of actin filaments in HeLa cells,
suggesting that phagocytosis is blocked via interference with the actin cyto-
skeleton. Jank and others went on to solve the structure of PaTox using
X-ray crystallography to a resolution of 1.8 A ( Jank et al., 2013). PaTox
is extremely interesting as it effectively performs two functions and the
crystal structure helps us understand how it performs both functions. First,
it targets Rho proteins (like TccC5 which is delivered by the Toxin Com-
plex Secretion System discussed above) by glycosylation of tyrosine
(GlcNAcylates) at Y32. Second, it also attacks heterotrimeric G proteins
via deamidation. Interestingly, P. asymbiotica does not carry a copy of TccC5,
and therefore, the authors only detected tyrosine GlcNAcylation during
P. asymbiotica infection of insect cells but not following infection with
P. luminescens. This might suggest that PaTox is somehow unique to, or
required by, Photorhabdus strains that can infect man but this hypothesis
remains to be tested.
The overall structure of the toxin is a single-chain AB toxin with two
catalytic domains within the C-terminus and a receptor-translocation
domain at the N-terminus ( Jank et al., 2013). Much like Mcf, its precise
mechanism of entry into the cell is uncertain, but it has been speculated that
it enters host cells from early endosomes ( Jank et al., 2013). The crystal
structure of PaToxG shows a unique GT-A fold, which is the smallest found
in currently known glycosyltransferase toxins, and lacks extending sub-
domains. As these sub-domains are critical in distinguishing Rho from
Ras as targets for enzymic attack, PaToxG is proposed to target Rho in a
highly specific manner. Downstream of the glycosyltransferase domain in
the crystal structure is a deamidase with amino acid similarity to Salmonella
SseI, and this PaTox domain is termed PaToxD ( Jank et al., 2013). Salmonella
SseI is secreted into host cells by the SPI-2 type III secretion system and is
thought to control the migration of host immune cells (McLaughlin et al.,
2009). The SseI-like domain of PaTox activates Gi and Gq/11 via
deamidation of a critical glutamine residue involved in the GTP hydrolysis
378 Richard H. ffrench-Constant and Andrea J. Dowling
6. BINARY TOXINS
Finally, we will briefly review the binary toxins that have been found
in either Photorhabdus or Xenorhabdus. These toxins are clearly smaller and
potentially easier to get to grips with than the large Photorhabdus toxins
and toxin delivery systems discussed so far. They may, therefore, also be eas-
ier to clone and utilize in insect control.
if a toxin effectively encoded its own secretion system (like the Tc secretion
system or TC-SS), then it should be possible to perform simple gain-of-
function screens to look for novel toxins active in recombinant E. coli.
We, therefore, designed a suite of simple gain-of-function screens that
we termed Rapid Virulence Annotation or RVA (Waterfield et al.,
2008). The first RVA-like screen obviously led to the discovery of Mcf1
and simply involved injecting individual recombinant E. coli clones carrying
large (3050 kb) insert sizes. As discussed above, this rapidly led to the dis-
covery of the dominant Mcf1 toxin which has then also served as a useful
positive control to test for the quality and insert size of any screened library.
Since Photorhabdus has to interact with, and suppress, a range of different
organisms, including its host nematode and a range of saprophytes such as
bacteria, nematodes and fungi that invade the infected insect corpse, we
expanded these simple screens to look at different insects, the model nem-
atode, C. elegans, amoebae (as models for phagocytosis), insect phagocytes
and macrophages (mammalian phagocytes). These screens are not only
highly effective in recovering loci that encode toxic proteins but also recov-
ered polyketide synthase and non-ribosomal peptide synthase encoding loci
that make a range of different small molecules and peptides (Waterfield et al.,
2008). Again, it is not the point of this review to simply provide a shopping
list of all the hits from these screens, which are reported elsewhere
(Waterfield et al., 2008), but to simply point out that the screens have rev-
ealed a further wealth of potentially bioactive proteins and small molecules
from Photorhabdus that demand further detailed investigation.
ACKNOWLEDGEMENTS
We would like to thank all members of the Richard ffrench-Constant laboratory, past and
present, for their work on Photorhabdus. We would also like to thank the BBSRC,
Leverhulme Trust and Royal Society of London for research funding.
REFERENCES
Ahantarig, A., Chantawat, N., Waterfield, N.R., ffrench-Constant, R., Kittayapong, P.,
2009. PirAB toxin from Photorhabdus asymbiotica as a larvicide against dengue vectors.
Appl. Environ. Microbiol. 75, 46274629.
Barth, H., Aktories, K., Popoff, M.R., Stiles, B.G., 2004. Binary bacterial toxins: biochem-
istry, biology, and applications of common Clostridium and Bacillus proteins. Microbiol.
Mol. Biol. Rev. 68, 373402 (table of contents).
384 Richard H. ffrench-Constant and Andrea J. Dowling
Blackburn, M., Golubeva, E., Bowen, D., ffrench-Constant, R.H., 1998. A novel insecti-
cidal toxin from Photorhabdus luminescens, Toxin complex a (Tca), and its histopatholog-
ical effects on the midgut of Manduca sexta. Appl. Environ. Microbiol. 64, 30363041.
Blackburn, M.B., Martin, P.A., Kuhar, D., Farrar Jr., R.R., Gundersen-Rindal, D.E., 2011.
The occurrence of Photorhabdus-like Toxin complexes in Bacillus thuringiensis. PLoS One
6, e18122.
Bowen, D.J., Ensign, J.C., 1998. Purification and characterization of a high-molecular-
weight insecticidal protein complex produced by the entomopathogenic bacterium Pho-
torhabdus luminescens. Appl. Environ. Microbiol. 64, 30293035.
Bowen, D., Rocheleau, T.A., Blackburn, M., Andreev, O., Golubeva, E., Bhartia, R.,
ffrench-Constant, R.H., 1998. Insecticidal toxins from the bacterium Photorhabdus
luminescens. Science 280, 21292132.
Brown, M.J., Thoren, K.L., Krantz, B.A., 2011. Charge requirements for proton gradient-
driven translocation of anthrax toxin. J. Biol. Chem. 286, 2318923199.
Brugirard-Ricaud, K., Duchaud, E., Givaudan, A., Girard, P.A., Kunst, F., Boemare, N.,
Brehelin, M., Zumbihl, R., 2005. Site-specific antiphagocytic function of the Photo-
rhabdus luminescens type III secretion system during insect colonization. Cell. Microbiol.
7, 363371.
Budd, R.C., 2001. Activation-induced cell death. Curr. Opin. Immunol. 13, 356362.
Busby, J.N., Panjikar, S., Landsberg, M.J., Hurst, M.R., Lott, J.S., 2013. The BC component
of ABC toxins is an RHS-repeat-containing protein encapsulation device. Nature
501, 547550.
Chaston, J.M., Suen, G., Tucker, S.L., Andersen, A.W., Bhasin, A., Bode, E., Bode, H.B.,
Brachmann, A.O., Cowles, C.E., Cowles, K.N., Darby, C., De Leon, L., Drace, K.,
Du, Z., Givaudan, A., Herbert Tran, E.E., Jewell, K.A., Knack, J.J., Krasomil-Osterfeld,
K.C., Kukor, R., Lanois, A., Latreille, P., Leimgruber, N.K., Lipke, C.M., Liu, R., Lu, X.,
Martens, E.C., Marri, P.R., Medigue, C., Menard, M.L., Miller, N.M., Morales-Soto, N.,
Norton, S., Ogier, J.C., Orchard, S.S., Park, D., Park, Y., Qurollo, B.A., Sugar, D.R.,
Richards, G.R., Rouy, Z., Slominski, B., Slominski, K., Snyder, H., Tjaden, B.C.,
Van Der Hoeven, R., Welch, R.D., Wheeler, C., Xiang, B., Barbazuk, B.,
Gaudriault, S., Goodner, B., Slater, S.C., Forst, S., Goldman, B.S., Goodrich-Blair, H.,
2011. The entomopathogenic bacterial endosymbionts Xenorhabdus and Photorhabdus: con-
vergent lifestyles from divergent genomes. PLoS One 6, e27909.
Chavez, C.V., Jubelin, G., Courties, G., Gomard, A., Ginibre, N., Pages, S., Taieb, F.,
Girard, P.A., Oswald, E., Givaudan, A., Zumbihl, R., Escoubas, J.M., 2010. The
cyclomodulin Cif of Photorhabdus luminescens inhibits insect cell proliferation and triggers
host cell death by apoptosis. Microbes Infect. 12, 12081218.
Choe, S., Bennett, M.J., Fujii, G., Curmi, P.M., Kantardjieff, K.A., Collier, R.J.,
Eisenberg, D., 1992. The crystal structure of diphtheria toxin. Nature 357, 216222.
Ciche, T.A., Ensign, J.C., 2003. For the insect pathogen Photorhabdus luminescens, which end
of a nematode is out? Appl. Environ. Microbiol. 69, 18901897.
Crow, A., Hughes, R.K., Taieb, F., Oswald, E., Banfield, M.J., 2012. The molecular basis of
ubiquitin-like protein NEDD8 deamidation by the bacterial effector protein Cif. Proc.
Natl. Acad. Sci. U.S.A. 109, E1830E1838.
Daborn, P.J., Waterfield, N., Blight, M.A., ffrench-Constant, R.H., 2001. Measuring vir-
ulence factor expression by the pathogenic bacterium Photorhabdus luminescens in culture
and during insect infection. J. Bacteriol. 183, 58345839.
Daborn, P.J., Waterfield, N., Silva, C.P., Au, C.P., Sharma, S., ffrench-Constant, R.H.,
2002. A single Photorhabdus gene, makes caterpillars floppy (mcf ), allows Escherichia coli
to persist within and kill insects. Proc. Natl. Acad. Sci. U.S.A. 99, 1074210747.
Dowling, A.J., Hodgson, D.J., 2014. An unbiased method for clustering bacterial effectors
using host cellular phenotypes. Appl. Environ. Microbiol. 80, 11851196.
Photorhabdus Toxins 385
Dowling, A.J., Daborn, P.J., Waterfield, N.R., Wang, P., Streuli, C.H., ffrench-Constant,
R.H., 2004. The insecticidal toxin Makes caterpillars floppy (Mcf ) promotes apoptosis in
mammalian cells. Cell. Microbiol. 6, 345353.
Duchaud, E., Rusniok, C., Frangeul, L., Buchrieser, C., Givaudan, A., Taourit, S., Bocs, S.,
Boursaux-Eude, C., Chandler, M., Charles, J.F., Dassa, E., Derose, R., Derzelle, S.,
Freyssinet, G., Gaudriault, S., Medigue, C., Lanois, A., Powell, K., Siguier, P.,
Vincent, R., Wingate, V., Zouine, M., Glaser, P., Boemare, N., Danchin, A.,
Kunst, F., 2003. The genome sequence of the entomopathogenic bacterium Photorhabdus
luminescens. Nat. Biotechnol. 21, 13071313.
ffrench-Constant, R.H., Waterfield, N., Burland, V., Perna, N.T., Daborn, P.J., Bowen, D.,
Blattner, F.R., 2000. A genomic sample sequence of the entomopathogenic bacterium
Photorhabdus luminescens W14: potential implications for virulence. Appl. Environ.
Microbiol. 66, 33103329.
ffrench-Constant, R., Waterfield, N., Daborn, P., Joyce, S., Bennett, H., Au, C.,
Dowling, A., Boundy, S., Reynolds, S., Clarke, D., 2003. Photorhabdus: towards a func-
tional genomic analysis of a symbiont and pathogen. FEMS Microbiol. Rev.
26, 433456.
ffrench-Constant, R.H., Dowling, A., Waterfield, N.R., 2007. Insecticidal toxins from Pho-
torhabdus bacteria and their potential use in agriculture. Toxicon 49, 436451.
Forst, S., Nealson, K., 1996. Molecular biology of the symbiotic-pathogenic bacteria
Xenorhabdus spp. and Photorhabdus spp.. Microbiol. Rev. 60, 2143.
Forst, S., Dowds, B., Boemare, N., Stackebrandt, E., 1997. Xenorhabdus and Photorhabdus
spp.: bugs that kill bugs. Annu. Rev. Microbiol. 51, 4772.
Gatsogiannis, C., Lang, A.E., Meusch, D., Pfaumann, V., Hofnagel, O., Benz, R.,
Aktories, K., Raunser, S., 2013. A syringe-like injection mechanism in Photorhabdus
luminescens toxins. Nature 495, 520523.
Gendlina, I., Held, K.G., Bartra, S.S., Gallis, B.M., Doneanu, C.E., Goodlett, D.R.,
Plano, G.V., Collins, C.M., 2007. Identification and type III-dependent secretion of
the Yersinia pestis insecticidal-like proteins. Mol. Microbiol. 64, 12141227.
Gerrard, J., Waterfield, N., Vohra, R., ffrench-Constant, R., 2004. Human infection with
Photorhabdus asymbiotica: an emerging bacterial pathogen. Microbes Infect. 6, 229237.
Gerrard, J.G., Joyce, S.A., Clarke, D.J., ffrench-Constant, R.H., Nimmo, G.R., Looke, D.F.,
Feil, E.J., Pearce, L., Waterfield, N.R., 2006. Nematode symbiont for Photorhabdus
asymbiotica. Emerg. Infect. Dis. 12, 15621564.
Heymann, J.B., Bartho, J.D., Rybakova, D., Venugopal, H.P., Winkler, D.C., Sen, A.,
Hurst, M.R., Mitra, A.K., 2013. Three-dimensional structure of the toxin-delivery par-
ticle antifeeding prophage of Serratia entomophila. J. Biol. Chem. 288, 2527625284.
Hong, W., Mosca, T.J., Luo, L., 2012. Teneurins instruct synaptic partner matching in an
olfactory map. Nature 484, 201207.
Huang, Y., Callahan, S., Hadfield, M.G., 2012. Recruitment in the sea: bacterial genes
required for inducing larval settlement in a polychaete worm. Sci. Rep. 2, 228.
Hurst, M.R., Glare, T.R., Jackson, T.A., Ronson, C.W., 2000. Plasmid-located pathoge-
nicity determinants of Serratia entomophila, the causal agent of amber disease of grass grub,
show similarity to the insecticidal toxins of Photorhabdus luminescens. J. Bacteriol.
182, 51275138.
Hurst, M.R., Jones, S.A., Binglin, T., Harper, L.A., Jackson, T.A., Glare, T.R., 2011. The
main virulence determinant of Yersinia entomophaga MH96 is a broad-host-range toxin
complex active against insects. J. Bacteriol. 193, 19661980.
Jank, T., Bogdanovic, X., Wirth, C., Haaf, E., Spoerner, M., Bohmer, K.E.,
Steinemann, M., Orth, J.H., Kalbitzer, H.R., Warscheid, B., Hunte, C.,
Aktories, K., 2013. A bacterial toxin catalyzing tyrosine glycosylation of Rho and
deamidation of Gq and Gi proteins. Nat. Struct. Mol. Biol. 20, 12731280.
386 Richard H. ffrench-Constant and Andrea J. Dowling
Kung, V.L., Khare, S., Stehlik, C., Bacon, E.M., Hughes, A.J., Hauser, A.R., 2012. An rhs
gene of Pseudomonas aeruginosa encodes a virulence protein that activates the
inflammasome. Proc. Natl. Acad. Sci. U.S.A. 109, 12751280.
Kupferschmied, P., Pechy-Tarr, M., Imperiali, N., Maurhofer, M., Keel, C., 2014. Domain
shuffling in a sensor protein contributed to the evolution of insect pathogenicity in plant-
beneficial Pseudomonas protegens. PLoS Pathog. 10, e1003964.
Landsberg, M.J., Jones, S.A., Rothnagel, R., Busby, J.N., Marshall, S.D., Simpson, R.M.,
Lott, J.S., Hankamer, B., Hurst, M.R., 2011. 3D structure of the Yersinia entomophaga
toxin complex and implications for insecticidal activity. Proc. Natl. Acad. Sci. U.S.A.
108, 2054420549.
Lang, A.E., Schmidt, G., Schlosser, A., Hey, T.D., Larrinua, I.M., Sheets, J.J., Mannherz, H.G.,
Aktories, K., 2010. Photorhabdus luminescens toxins ADP-ribosylate actin and RhoA to force
actin clustering. Science 327, 11391142.
Lang, A.E., Schmidt, G., Sheets, J.J., Aktories, K., 2011. Targeting of the actin cytoskeleton
by insecticidal toxins from Photorhabdus luminescens. Naunyn Schmiedebergs Arch.
Pharmacol. 383, 227235.
Lang, A.E., Konukiewitz, J., Aktories, K., Benz, R., 2013. TcdA1 of Photorhabdus
luminescens: electrophysiological analysis of pore formation and effector binding.
Biophys. J. 105, 376384.
Lang, A.E., Ernst, K., Lee, H., Papatheodorou, P., Schwan, C., Barth, H., Aktories, K.,
2014. The chaperone Hsp90 and PPIases of the cyclophilin and FKBP families facilitate
membrane translocation of Photorhabdus luminescens ADP-ribosyltransferases. Cell.
Microbiol. 16, 490503.
Lanois, A., Ogier, J.C., Gouzy, J., Laroui, C., Rouy, Z., Givaudan, A., Gaudriault, S., 2013.
Draft genome sequence and annotation of the entomopathogenic bacterium Xenorhabdus
nematophila strain F1. Genome Announc. 1 e00342-13.
Lee, S.C., Stoilova-Mcphie, S., Baxter, L., Fulop, V., Henderson, J., Rodger, A.,
Roper, D.I., Scott, D.J., Smith, C.J., Morgan, J.A., 2007. Structural characterisation
of the insecticidal toxin XptA1, reveals a 1.15 MDa tetramer with a cage-like structure.
J. Mol. Biol. 366, 15581568.
Leiman, P.G., Basler, M., Ramagopal, U.A., Bonanno, J.B., Sauder, J.M., Pukatzki, S.,
Burley, S.K., Almo, S.C., Mekalanos, J.J., 2009. Type VI secretion apparatus and phage
tail-associated protein complexes share a common evolutionary origin. Proc. Natl. Acad.
Sci. U.S.A. 106, 41544159.
Liu, D., Burton, S., Glancy, T., Li, Z.S., Hampton, R., Meade, T., Merlo, D.J., 2003. Insect
resistance conferred by 283-kDa Photorhabdus luminescens protein TcdA in Arabidopsis
thaliana. Nat. Biotechnol. 21, 12221228.
Marshall, S.D., Hares, M.C., Jones, S.A., Harper, L.A., Vernon, J.R., Harland, D.P.,
Jackson, T.A., Hurst, M.R., 2012. Histopathological effects of the Yen-Tc toxin
complex from Yersinia entomophaga MH96 (Enterobacteriaceae) on the Costelytra
zealandica (Coleoptera: Scarabaeidae) larval midgut. Appl. Environ. Microbiol. 78,
48354847.
Mclaughlin, L.M., Govoni, G.R., Gerke, C., Gopinath, S., Peng, K., Laidlaw, G., Chien, Y.H.,
Jeong, H.W., Li, Z., Brown, M.D., Sacks, D.B., Monack, D., 2009. The Salmonella SPI2
effector SseI mediates long-term systemic infection by modulating host cell migration. PLoS
Pathog. 5, e1000671.
Meusch, D., Gatsogiannis, C., Efremov, R.G., Lang, A.E., Hofnagel, O., Vetter, I.R.,
Aktories, K., Raunser, S., 2014. Mechanism of Tc toxin action revealed in molecular
detail. Nature 508, 6165.
Moak, M., Molineux, I.J., 2004. Peptidoglycan hydrolytic activities associated with bacte-
riophage virions. Mol. Microbiol. 51, 11691183.
Photorhabdus Toxins 387
Morgan, J.A., Sergeant, M., Ellis, D., Ousley, M., Jarrett, P., 2001. Sequence analysis of
insecticidal genes from Xenorhabdus nematophilus PMFI296. Appl. Environ. Microbiol.
67, 20622069.
Nakayama, K., Takashima, K., Ishihara, H., Shinomiya, T., Kageyama, M., Kanaya, S.,
Ohnishi, M., Murata, T., Mori, H., Hayashi, T., 2000. The R-type pyocin of Pseudo-
monas aeruginosa is related to P2 phage, and the F-type is related to lambda phage. Mol.
Microbiol. 38, 213231.
Peat, S.M., ffrench-Constant, R.H., Waterfield, N.R., Marokhazi, J., Fodor, A., Adams, B.J.,
2010. A robust phylogenetic framework for the bacterial genus Photorhabdus and its use in
studying the evolution and maintenance of bioluminescence: a case for 16S, gyrB, and
glnA. Mol. Phylogenet. Evol. 57, 728740.
Pechy-Tarr, M., Bruck, D.J., Maurhofer, M., Fischer, E., Vogne, C., Henkels, M.D.,
Donahue, K.M., Grunder, J., Loper, J.E., Keel, C., 2008. Molecular analysis of a novel
gene cluster encoding an insect toxin in plant-associated strains of Pseudomonas fluorescens.
Environ. Microbiol. 10, 23682386.
Pechy-Tarr, M., Borel, N., Kupferschmied, P., Turner, V., Binggeli, O., Radovanovic, D.,
Maurhofer, M., Keel, C., 2013. Control and host-dependent activation of insect toxin expres-
sion in a root-associated biocontrol pseudomonad. Environ. Microbiol. 15, 736750.
Petosa, C., Collier, R.J., Klimpel, K.R., Leppla, S.H., Liddington, R.C., 1997. Crystal struc-
ture of the anthrax toxin protective antigen. Nature 385, 833838.
Purcell, J.P., Greenplate, J.T., Jennings, M.G., Ryerse, J.S., Pershing, J.C., Sims, S.R.,
Prinsen, M.J., Corbin, D.R., Tran, M., Sammons, R.D., et al., 1993. Cholesterol oxi-
dase: a potent insecticidal protein active against boll weevil larvae. Biochem. Biophys.
Res. Commun. 196, 14061413.
Rodou, A., Ankrah, D.O., Stathopoulos, C., 2010. Toxins and secretion systems of Photo-
rhabdus luminescens. Toxins (Basel) 2, 12501264.
Sergeant, M., Jarrett, P., Ousley, M., Morgan, J.A., 2003. Interactions of insecticidal toxin
gene products from Xenorhabdus nematophilus PMFI296. Appl. Environ. Microbiol.
69, 33443349.
Sheets, J.J., Hey, T.D., Fencil, K.J., Burton, S.L., Ni, W., Lang, A.E., Benz, R., Aktories, K.,
2011. Insecticidal toxin complex proteins from Xenorhabdus nematophilus: structure and
pore formation. J. Biol. Chem. 286, 2274222749.
Shikuma, N.J., Pilhofer, M., Weiss, G.L., Hadfield, M.G., Jensen, G.J., Newman, D.K.,
2014. Marine tubeworm metamorphosis induced by arrays of bacterial phage tail-like
structures. Science 343, 529533.
Somvanshi, V.S., Sloup, R.E., Crawford, J.M., Martin, A.R., Heidt, A.J., Kim, K.S.,
Clardy, J., Ciche, T.A., 2012. A single promoter inversion switches Photorhabdus
between pathogenic and mutualistic states. Science 337, 8893.
Spinner, J.L., Jarrett, C.O., Larock, D.L., Miller, S.I., Collins, C.M., Hinnebusch, B.J., 2012.
Yersinia pestis insecticidal-like Toxin complex (Tc) family proteins: characterization of
expression, subcellular localization, and potential role in infection of the flea vector.
BMC Microbiol. 12, 296.
Tan, B., Jackson, T.A., Hurst, M.R., 2006. Virulence of Serratia strains against Costelytra
zealandica. Appl. Environ. Microbiol. 72, 64176418.
Vadyvaloo, V., Jarrett, C., Sturdevant, D.E., Sebbane, F., Hinnebusch, B.J., 2010. Transit
through the flea vector induces a pretransmission innate immunity resistance phenotype
in Yersinia pestis. PLoS Pathog. 6, e1000783.
Vigneux, F., Zumbihl, R., Jubelin, G., Ribeiro, C., Poncet, J., Baghdiguian, S.,
Givaudan, A., Brehelin, M., 2007. The xaxAB genes encoding a new apoptotic toxin
from the insect pathogen Xenorhabdus nematophila are present in plant and human path-
ogens. J. Biol. Chem. 282, 95719580.
388 Richard H. ffrench-Constant and Andrea J. Dowling
Visschedyk, D.D., Perieteanu, A.A., Turgeon, Z.J., Fieldhouse, R.J., Dawson, J.F.,
Merrill, A.R., 2010. Photox, a novel actin-targeting mono-ADP-ribosyltransferase from
Photorhabdus luminescens. J. Biol. Chem. 285, 1352513534.
Vlisidou, I., Dowling, A.J., Evans, I.R., Waterfield, N., ffrench-Constant, R.H., Wood, W.,
2009. Drosophila embryos as model systems for monitoring bacterial infection in real
time. PLoS Pathog. 5, e1000518.
Wagner, R., Wolff, T., Herwig, A., Pleschka, S., Klenk, H.D., 2000. Interdependence of
hemagglutinin glycosylation and neuraminidase as regulators of influenza virus growth:
a study by reverse genetics. J. Virol. 74, 63166323.
Wagner, N.J., Lin, C.P., Borst, L.B., Miller, V.L., 2013. YaxAB, a Yersinia enterocolitica pore-
forming toxin regulated by RovA. Infect. Immun. 81, 42084219.
Waterfield, N., Dowling, A., Sharma, S., Daborn, P.J., Potter, U., ffrench-Constant, R.H.,
2001a. Oral toxicity of Photorhabdus luminescens W14 toxin complexes in Escherichia coli.
Appl. Environ. Microbiol. 67, 50175024.
Waterfield, N.R., Bowen, D.J., Fetherston, J.D., Perry, R.D., ffrench-Constant, R.H.,
2001b. The tc genes of Photorhabdus: a growing family. Trends Microbiol. 9, 185191.
Waterfield, N.R., Daborn, P.J., Dowling, A.J., Yang, G., Hares, M., ffrench-Constant, R.H.,
2003. The insecticidal toxin Makes caterpillars floppy 2 (Mcf2) shows similarity to HrmA,
an avirulence protein from a plant pathogen. FEMS Microbiol. Lett. 229, 265270.
Waterfield, N., Kamita, S.G., Hammock, B.D., ffrench-Constant, R., 2005. The Photo-
rhabdus Pir toxins are similar to a developmentally regulated insect protein but show
no juvenile hormone esterase activity. FEMS Microbiol. Lett. 245, 4752.
Waterfield, N.R., Sanchez-Contreras, M., Eleftherianos, I., Dowling, A., Yang, G.,
Wilkinson, P., Parkhill, J., Thomson, N., Reynolds, S.E., Bode, H.B., Dorus, S.,
ffrench-Constant, R.H., 2008. Rapid Virulence Annotation (RVA): identification of
virulence factors using a bacterial genome library and multiple invertebrate hosts. Proc.
Natl. Acad. Sci. U.S.A. 105, 1596715972.
Waterfield, N.R., Ciche, T., Clarke, D., 2009. Photorhabdus and a host of hosts. Annu. Rev.
Microbiol. 63, 557574.
Wiggins, C.A., Munro, S., 1998. Activity of the yeast MNN1 alpha-1,3-mannosyltransferase
requires a motif conserved in many other families of glycosyltransferases. Proc. Natl.
Acad. Sci. U.S.A. 95, 79457950.
Wilkinson, P., Waterfield, N.R., Crossman, L., Corton, C., Sanchez-Contreras, M.,
Vlisidou, I., Barron, A., Bignell, A., Clark, L., Ormond, D., Mayho, M., Bason, N.,
Smith, F., Simmonds, M., Churcher, C., Harris, D., Thompson, N.R., Quail, M.,
Parkhill, J., ffrench-Constant, R.H., 2009. Comparative genomics of the emerging
human pathogen Photorhabdus asymbiotica with the insect pathogen Photorhabdus
luminescens. BMC Genomics 10, 302.
Wilkinson, P., Paszkiewicz, K., Moorhouse, A., Szubert, J.M., Beatson, S., Gerrard, J.,
Waterfield, N.R., ffrench-Constant, R.H., 2010. New plasmids and putative virulence
factors from the draft genome of an Australian clinical isolate of Photorhabdus asymbiotica.
FEMS Microbiol. Lett. 309, 136143.
Yang, G., Waterfield, N.R., 2013. The role of TcdB and TccC subunits in secretion of the
Photorhabdus Tcd toxin complex. PLoS Pathog. 9, e1003644.
Yang, G., Dowling, A.J., Gerike, U., ffrench-Constant, R.H., Waterfield, N.R., 2006. Pho-
torhabdus virulence cassettes confer injectable insecticidal activity against the wax moth.
J. Bacteriol. 188, 22542261.
Zhang, D., De Souza, R.F., Anantharaman, V., Iyer, L.M., Aravind, L., 2012. Polymorphic
toxin systems: comprehensive characterization of trafficking modes, processing, mech-
anisms of action, immunity and ecology using comparative genomics. Biol. Direct 7, 18.
Zhang, X., Hu, X., Li, Y., Ding, X., Yang, Q., Sun, Y., Yu, Z., Xia, L., Hu, S., 2014.
XaxAB-like binary toxin from Photorhabdus luminescens exhibits both insecticidal activity
and cytotoxicity. FEMS Microbiol. Lett. 350, 4856.
CHAPTER EIGHT
Contents
1. Introduction 390
2. Spider Venom Peptides as Bioinsecticides 391
3. Transcytosis of Spider Venom Peptides Across the Insect Gut Epithelium
via Fusion to Molecular Transport Vehicles 394
4. Use of Entomopathogens for ISVP Delivery 397
4.1 Entomopathogens as bioinsecticides 397
4.2 Entomopathogens as a delivery vehicle for ISVPs 398
5. In Planta Expression of Spider Venom Peptides 399
6. ISVP Mimetics 401
6.1 Synthetic mimics of venom peptides 401
6.2 Hv1a pharmacophore analysis and modelling 402
6.3 Optimisation and mechanism of action of the Type II mimic VNX-440 405
6.4 Conclusions from development of Type-II mimetics of Hv1a 407
7. Outlook 407
Acknowledgements 408
References 408
Abstract
Despite intensive control measures, insect pests cause enormous damage to crops and
stored grain. In addition, insects vector some of the world's most devastating human
diseases, including malaria, dengue and Chagas disease. Chemical insecticides remain
the dominant method of controlling insect pests but the arsenal of extant insecticides is
rapidly diminishing due to the evolution of resistance in pest species and a more dif-
ficult regulatory environment due to heightened concern about the potential adverse
impacts of chemical insecticides on human health and the environment. Along with
predatory beetles, spiders are the most successful insect predators on the planet and
their venoms contain a diverse array of small, disulfide-rich insecticidal peptides. In addi-
tion to being potent and highly selective for insects, they collectively offer very diverse
pharmacology and should degrade to innocuous breakdown products in the field.
However, their major disadvantage is a low level of intrinsic oral activity. Consequently,
much research has been directed towards improving the oral activity of these peptide
toxins or, alternatively, obviating the oral activity problem altogether by incorporating
transgenes encoding the toxins into suitable biological delivery vehicles. Here, we dis-
cuss recent advances in both of these areas. We also discuss an approach that merges
the advantages of small-molecule and peptide-based insecticides, namely using the
pharmacophore of an insecticidal spider-venom peptide to rationally develop a
small-molecule mimetic that has improved oral activity, but retains activity for a novel
insecticide target.
1. INTRODUCTION
The human population is projected to reach to 9.3 billion by 2050
(Bloom, 2011). This expansion of the population will create enormous pres-
sure on food production since there is limited scope for increasing the
amount of cultivated land (Oerke, 2006). Despite intensive control mea-
sures, herbivorous insects reduce world crop yields by 1014% (Oerke,
2006) and damage as much as 30% of stored grain (Boyer et al., 2012),
and hence their control is a key factor in determining food productivity.
Increases in human population density combined with global warming
and a dramatic escalation in international travel has also placed considerable
pressure on the public health sector to control the spread of insects that vec-
tor human diseases such as malaria, dengue and Chagas disease.
Unfortunately, the arsenal of chemical insecticides available to combat
insect pests is in rapid decline. This is largely due to the fact that most chem-
ical insecticides target a handful of enzymes, receptors and ion channels in
the insect nervous system, and consequently their long-term use has led to
widespread resistance in pest species (King and Hardy, 2013; Smith et al.,
2013). Concerns about the adverse impact of certain classes of chemical
insecticides on the environment and human health have also led to wide-
spread de-registrations and use cancellations; for example, over the period
20052009, the U.S. Environmental Protection Agency (EPA)
de-registered or limited the use of 169 insecticides, while only nine new
insecticides were registered during the same time period (Windley et al.,
2012). In simple terms, we have to combat an increasing insect pest problem
with a rapidly diminishing chemical arsenal.
Spider Venom Peptides as Bioinsecticides 391
The biggest potential disadvantage of ISVPs is that they are not orally or
topically active. Spiders have solved the delivery problem by evolving fangs
that act like hypodermic syringes to deliver their venoms into the
haemolymph of prey, thus enabling ISVPs to access their targets in the insect
nervous system. We have found that the LD50 values for ISVPs fed to insects
are about 100-fold higher than when they are injected (Hardy et al., 2013;
King and Hardy, 2013), presumably due to a slow rate of absorption in
the insect gut, as observed previously for insect neuropeptides (Audsley
et al., 2008) and disulfide-rich peptides from scorpion and snake venoms
(Casartelli et al., 2005). Nevertheless, the low level of intrinsic oral activity
of ISVPs is not restricted to insects, as other arthropod pests such as ticks
were found to be sensitive to oral administration of -hexatoxin-Hv1a
(hereafter Hv1a) (Mukherjee et al., 2006).
There is growing evidence that some peptides and proteins can access
paracellular and transcellular transport routes in the insect midgut. Septate
junctions in insects are leaky for molecules smaller than 5 kDa (Casartelli
et al., 2005; Skaer et al., 1987), and this paracellular route might explain
the observed uptake of insect neuropeptides and ISVPs (Audsley et al.,
2007; Hardy et al., 2013). However, size alone is not the sole determinant
of uptake rate, as the 0.9-kDa insect neuropeptide cydiastatin-4 is translocated
across the midgut of the lepidopteran Manduca sexta at a rate 42 times lower
than for Hv1a, a 4-kDa ISVP (V. Herzig, N. Audsley, and G.F. King,
unpublished data). In addition to the paracellular route, the translocation of
large proteins such as albumin across the insect midgut epithelium involves
active energy-dependent processes such as transcytosis (Casartelli et al., 2005).
Even though our understanding of transport routes across the insect mid-
gut is still in its infancy, it is clear that some peptides can pass undegraded
through the midgut, consistent with the observed insecticidal effects of ISVPs
after oral administration. Further research in this area might facilitate the
rational introduction of chemical modifications and/or the use of formula-
tions that will enhance the oral activity of ISVPs and facilitate their direct
deployment as bioinsecticides. However, the remainder of this chapter will
focus on alternative methods of delivering ISVPs to target insect species as
well as the option of developing orally active small-molecule ISVP mimetics.
Figure 8.1 Summary of methods for delivering ISVPs to the insect haemocoel: (1) Some
ISVPs appear to be able to access the haemocoel via paracellular uptake through leaky
septate junctions. (2) ISVPs can be fused to carrier proteins such as snowdrop lectin
(GNA) and the coat protein of pea enation mosaic virus, which are recognised by specific
receptors on gut epithelial cells. These fusion proteins can be taken up via receptor-
mediated endocytosis and delivered to the haemocoel by transcytosis. (3) Baculovirus
can be engineered to encode an ISVP transgene. Upon recognition by specific gut
receptors, budded virions infect gut epithelial cells and other body tissues and release
ISVPs. (4) Entomopathogenic fungi such as Metarhizium and Beauveria can also be
engineered to encode an ISVP transgene. Fungal spores (conidia) adhere to the insect
cuticle, germinate and then invade the integument. The fungus then proliferates in the
haemocoel in the form of wall-less blastopores, which produce and release ISVPs.
Adapted from Bonning and Chougule (2014).
during which time the insect continues to feed and cause crop damage
(Bonning and Hammock, 1996; Inceoglu et al., 2006).
There are 400500 species of entomopathogenic fungi and some have
been used, without modification, for biocontrol of pest insects for 150 years
(Bailey and Boyetchko, 2010; Shah and Pell, 2003). Compared to other
micro-organisms, fungi infect a broader range of insects including lepidop-
terans, homopterans, hymenopterans, coleopterans and dipterans (Vilcinskas
et al., 1997). The EPA have registered more than 100 fungus-based insect
control products since 1995 (De Faria and Wraight, 2007; Shah and Pell,
2003). Mycotrol, for example, is a Beauveria-based mycoinsecticide used
for control of whiteflies, thrips, mealybugs and aphids (De Faria and
Wraight, 2007), while Green Muscle contains spores of Metarhizium
acridum, which are selectively pathogenic to locusts. Green Muscle showed
no detrimental effects on non-target organisms and it is now registered and
recommended for locust control by the United Nations (Lomer et al., 2001).
However, as for baculoviruses, a major disadvantage of entomopathogenic
fungi compared with chemical insecticides is their slow kill time (typically
>7 days).
to the introduction of Bt crops have now become a major pest on some crops
(Bonning and Chougule, 2014).
Resistance to Bt crops can be delayed through the use of non-GM ref-
uges and/or by engineering Bt plants to express additional insecticidal genes
that act via different mechanisms, an approach known as pyramiding or trait
stacking (Moar and Anilkumar, 2007; Que et al., 2010; Chapter 6). ISVP
transgenes are good candidates for trait stacking with Bt since (i) they have
completely different mechanisms of action; (ii) they are likely to be
synergised by Cry toxins which lyse midgut epithelial cells (Soberon
et al., 2007), a property that should facilitate translocation of ISVPs into
the insect haemocoel; and (iii) whereas Bt toxins are largely specific for
the insect orders Lepidoptera and Coleoptera, ISVPs with complementary
selectivity, particularly against sap-sucking hemipterans, can be selected
for trait stacking.
Attempts to engineer plants expressing ISVPs began almost two decades
ago with the demonstration that transgenic tobacco expressing -HXTX-
Ar1a, a 37-residue insect-specific calcium channel blocker from the
Australian funnel-web spider, Atrax robustus, had enhanced resistance to
H. armigera ( Jiang et al., 1996). Transgenes encoding this ISVP (or its
orthologue Hv1a) have subsequently been engineered into cotton (Omar
and Chatha, 2012), tobacco (Khan et al., 2006) and poplar (Cao et al.,
2010). All of these ISVP-expressing transgenic plants were shown to have
significantly increased resistance to insect pests. For example, the mortality
of second instar H. armigera fed on transgenic tobacco expressing Hv1a was
75100% after 72 h compared to 0% for larvae fed on untransformed plants,
regardless of whether the ISVP was expressed under the strong 35S pro-
moter (Khan et al., 2006) or weaker phloem-specific promoters (Shah
et al., 2011). It has even been claimed that transgenic cotton expressing
Hv1a is as effective as Bollgard II cotton in controlling major cotton pests
(Omar and Chatha, 2012).
Tobacco plants engineered to express U7-HXTX-Mg1a (Magi-6), a
36-residue ISVP from the venom of the hexathelid spider Macrothele gigas
were found to be significantly more resistant than wild-type plants
against S. frugiperda (Hernandez-Campuzano et al., 2009). More recently,
transgenic Arabidopsis expressing Hv1a fused to the CP of PEMV were
shown to be resistant to a diverse range of aphid species (Bonning et al.,
2014). Thus, ISVP transgenes appear to have significant potential
either as a standalone insect-resistant plant trait or for trait stacking with
Cry toxins.
Spider Venom Peptides as Bioinsecticides 401
6. ISVP MIMETICS
6.1. Synthetic mimics of venom peptides
Venom peptides present structural characteristics that are far from the norm
for synthetic agrochemicals and drugs (Tice, 2001), and their size and polar-
ity impede passive diffusion through cell membranes. Instead, their oral bio-
availability to targets beyond the gut lumen relies on paracellular or active
transport (see Section 2). Given the diversity and potential complexity of
mechanisms that can mediate such paracellular/active transit through the
physical barriers of the gut wall, it seems unlikely that a simple set of general
rules could be derived that describe the oral bioavailability of venom pep-
tides. While the oral activity of ISVPs might be improved by chemical mod-
ifications (e.g. PEGylation and acylation) or by packaging them in
liposomes, microparticles and/or nanoparticles, these are complex and
expensive modifications that would significantly increase production costs.
An alternative approach to improving oral activity is to create
peptidomimetics wherein the functionally critical chemical features of the
peptide are displayed on a smaller structural scaffold with better pharmaco-
kinetic properties.
This alternative has been most extensively pursued for the conotoxin
family of venom peptides. The physicochemical properties of Ziconotide/
Prialt (-conotoxin MVIIA), a cone snail venom peptide used for manage-
ment of intractable pain, result in insignificant oral bioavailability and
consequently this drug requires intrathecal administration (Bingham et al.,
2010). In order to avoid the complexity of this drug delivery method, efforts
have been undertaken to design small-molecule mimetics that might show
oral bioavailability and bloodbrain barrier permeability (Brady et al., 2013).
Peptidomimetics that could surmount these challenges are classifiable
into three types (Ripka and Rich, 1998): (Type I) molecules in which pep-
tide bonds are replaced with more tractable bioisosteres; (Type II) molecules
which use moieties that are not structurally equivalent to the native peptide
to interact with the receptor; and (Type III) molecules which utilise a scaf-
fold that can spatially display the critical amino acid residues in the same way
as the native peptide. Several papers have been published on the design of
Type-III mimetics of -conotoxins due to the medical and commercial
need (Baell et al., 2001, 2004; Menzler et al., 2000). Despite early progress,
the resultant conotoxin mimetics were much less potent than the native pep-
tide (Brady et al., 2013).
402 Volker Herzig et al.
Available SAR data suggested that the most functionally important res-
idue of Hv1a is Arg35 (Tedford et al., 2004a). This residue contributed three
features to the pharmacophore models: two hydrogen-bond donors (HBDs)
and one positively ionisable group. Asn27 contributed two features: one
HBD and one HBA. Pro10 and Val33 presumably mediate hydrophobic
interactions with CaV channels targeted by the peptide; hence, hydrophobic
features represented their side chains in the pharmacophore. Gly8, which
adopts an unusual backbone configuration (backbone dihedral angles:
90 and 20 ), is conformationally restricted by the cystine knot
in a way that allows the Gly8 carbonyl oxygen to contribute to the toxin
pharmacophore.
404 Volker Herzig et al.
Table 8.2 Chemical features of key Hv1a residues used to build 3D pharmacophore
models
Arg35 Asn27 Val33 Pro10 Gly8
Model HBD HBD PI HBA HBD HPH (C, HPH (C, HBA
version (N1) (N2) (N) (O1) (N2) C1, C2) C, C) (O)
8-point Y Y Y Y Y Y Y Y
6-point 1 Y Y Y Y Y N Y N
6-point 2 N Y Y Y Y Y Y N
HBD, hydrogen-bond donor; PI, positively ionisable group; HBA, hydrogen-bond acceptor; HPH,
hydrophobe.
Spider Venom Peptides as Bioinsecticides 405
the cationic feature; and, finally, the remaining anilide has modest overlap
with the two hydrophobic features.
7. OUTLOOK
An unmodified ISVP developed by Vestaron Corporation was
recently approved by the U.S. EPA for use against cabbage looper, Tricho-
plusia ni, marking the start of new era of peptide-based insecticides. ISVPs
have many properties that make them attractive as insect control agents: they
are potently insecticidal, highly selective for insects, stable to extremes of
temperature and should degrade in the field into innocuous breakdown
products. Although they typically have low levels of intrinsic oral activity,
simple formulation can improve their oral activity sufficiently to make them
competitive with chemical insecticides. The oral activity of ISVPs can also
be markedly improved by fusing them to proteins that ferry them via
transcytosis across the insect gut, including plant lectins and the CP of
insect-vectored viruses. Transgenes encoding these fusion proteins could
be incorporated into crop plants either as a standalone insect-resistance trait
or they could be pyramided with Bt transgenes in order to reduce the like-
lihood of resistance development and expand the range of susceptible
insects. Entomopathogens can also be engineered to express transgenes
encoding ISVPs or ISVP fusion proteins with enhanced oral activity. This
approach has the advantage that the selectivity of the ISVP can be tailored
by taking advantage of the limited host range of many entomopathogens.
408 Volker Herzig et al.
Thus, in summary, there are a wide variety of methods by which ISVPs can
be deployed as natural insecticides. We also recently demonstrated for the
first time that the pharmacophore of ISVPs can be used to rationally develop
small-molecule mimetics with improved oral activity, thereby providing
another approach by which these peptides can be exploited for the control
of insect pests.
ACKNOWLEDGEMENTS
The authors would like to thank the Australian Research Council for financial support. R.M.
Kennedy was funded in part by a grant from the Foundation for the National Institutes of
Health through the Vector-based Transmission of Control: Discovery Research (VCTR)
program of the Grand Challenges in Global Health initiative.
REFERENCES
Atkinson, R.K., Howden, M.E.H., Tyler, M.I., Vonarx, E.J., 1998. Insecticidal toxins
derived from funnel web (Atrax or Hadronyche) spiders. U.S. Patent No. 5,763,568.
Audsley, N., Matthews, J., Nachman, R., Weaver, R.J., 2007. Metabolism of cydiastatin 4
and analogues by enzymes associated with the midgut and haemolymph of Manduca sexta
larvae. Gen. Comp. Endocrinol. 153, 8087.
Audsley, N., Matthews, J., Nachman, R.J., Weaver, R.J., 2008. Transepithelial flux of an
allatostatin and analogs across the anterior midgut of Manduca sexta larvae in vitro.
Peptides 29, 286294.
Baell, J.B., Forsyth, S.A., Gable, R.W., Norton, R.S., Mulder, R.J., 2001. Design and syn-
thesis of type-III mimetics of -conotoxin GVIA. J. Comput. Aided Mol. Des.
15, 11191136.
Baell, J.B., Duggan, P.J., Forsyth, S.A., Lewis, R.J., Lok, Y.P., Schroeder, C.I., 2004. Syn-
thesis and biological evaluation of nonpeptide mimetics of -conotoxin GVIA. Bioorg.
Med. Chem. 12, 40254037.
Bailey, K.L., Boyetchko, S.M., 2010. Social and economic drivers shaping the future of bio-
logical control: a Canadian perspective on the factors affecting the development and use
of microbial biopesticides. Biol. Control 52, 221229.
Bingham, J.P., Mitsunaga, E., Bergeron, Z.L., 2010. Drugs from slugspast, present and
future perspectives of conotoxin research. Chem. Biol. Interact. 183, 118.
Blackledge, T.A., Scharff, N., Coddington, J.A., Szuts, T., Wenzel, J.W., Hayashi, C.Y.,
Agnarsson, I., 2009. Reconstructing web evolution and spider diversification in the
molecular era. Proc. Natl. Acad. Sci. U.S.A. 106, 52295234.
Bloom, D.E., 2011. 7 billion and counting. Science 333, 562569.
Bonning, B.C., Chougule, N.P., 2014. Delivery of intrahemocoelic peptides for insect pest
management. Trends Biotechnol. 32, 9198.
Bonning, B.C., Hammock, B.D., 1996. Development of recombinant baculoviruses for
insect control. Annu. Rev. Entomol. 41, 191210.
Bonning, B.C., Pal, N., Liu, S., Wang, Z., Sivakumar, S., Dixon, P.M., King, G.F.,
Miller, W.A., 2014. Toxin delivery by the coat protein of an aphid-vectored plant virus
provides plant resistance to aphids. Nat. Biotechnol. 32, 102105.
Boyer, S., Zhang, H., Lemperiere, G., 2012. A review of control methods and resistance
mechanisms in stored-product insects. Bull. Entomol. Res. 102, 213229.
Brady, R.M., Baell, J.B., Norton, R.S., 2013. Strategies for the development of conotoxins
as new therapeutic leads. Mar. Drugs 11, 22932313.
Spider Venom Peptides as Bioinsecticides 409
Cao, C.W., Liu, G.F., Wang, Z.Y., Yan, S.C., Ma, L., Yang, C.P., 2010. Response of the
gypsy moth, Lymantria dispar to transgenic poplar, Populus simonii P. nigra, expressing
fusion protein gene of the spider insecticidal peptide and Bt-toxin C-peptide.
J. Insect Sci. 10, 200.
Casartelli, M., Corti, P., Giovanna Leonardi, M., Fiandra, L., Burlini, N., Pennacchio, F.,
Giordana, B., 2005. Absorption of albumin by the midgut of a lepidopteran larva.
J. Insect Physiol. 51, 933940.
Chong, Y., Hayes, J.L., Sollod, B., Wen, S., Wilson, D.T., Hains, P.G., Hodgson, W.C.,
Broady, K.W., King, G.F., Nicholson, G.M., 2007. The -atracotoxins: selective blockers
of insect M-LVA and HVA calcium channels. Biochem. Pharmacol. 74, 623638.
De Faria, M.R., Wraight, S.P., 2007. Mycoinsecticides and mycoacaricides, a comprehen-
sive list with worldwide coverage and international classification of formulation types.
Biol. Control 43, 237256.
Escoubas, P., Sollod, B., King, G.F., 2006. Venom landscapes: mining the complexity of spi-
der venoms via a combined cDNA and mass spectrometric approach. Toxicon
47, 650663.
Fitches, E., Woodhouse, S.D., Edwards, J.P., Gatehouse, J.A., 2001. In vitro and in vivo bind-
ing of snowdrop (Galanthus nivalis agglutinin; GNA) and jackbean (Canavalia ensiformis;
Con A) lectins within tomato moth (Lacanobia oleracea) larvae; mechanisms of insecticidal
action. J. Insect Physiol. 47, 777787.
Fitches, E., Edwards, M., Mee, C., Grishin, E., Gatehouse, A., Edwards, J., Gatehouse, J.,
2004. Fusion proteins containing insect-specific toxins as pest control agents: snowdrop
lectin delivers fused insecticidal spider venom toxin to insect haemolymph following oral
ingestion. J. Insect Physiol. 50, 6171.
Fitches, E.C., Pyati, P., King, G.F., Gatehouse, J.A., 2012. Fusion to snowdrop lectin mag-
nifies the oral activity of insecticidal -hexatoxin-Hv1a peptide by enabling its delivery
to the central nervous system. PLoS One 7, e39389.
Fletcher, J.I., Smith, R., Odonoghue, S.I., Nilges, M., Connor, M., Howden, M.E.,
Christie, M.J., King, G.F., 1997. The structure of a novel insecticidal neurotoxin,
-atracotoxin-HV1, from the venom of an Australian funnel web spider. Nat. Struct.
Biol. 4, 559566.
Gassmann, A.J., Petzold-Maxwell, J.L., Clifton, E.H., Dunbar, M.W., Hoffmann, A.M.,
Ingber, D.A., Keweshan, R.S., 2014. Field-evolved resistance by western corn
rootworm to multiple Bacillus thuringiensis toxins in transgenic maize. Proc. Natl. Acad.
Sci. U.S.A. 111, 51415146.
Gatehouse, A.M.R., Ferry, N., Edwards, M.G., Bell, H.A., 2011. Insect-resistant biotech
crops and their impacts on beneficial arthropods. Philos. Trans. R. Soc. Lond. B Biol.
Sci. 366, 14381452.
Hardy, M.C., Daly, N.L., Mobli, M., Morales, R.A., King, G.F., 2013. Isolation of an orally
active insecticidal toxin from the venom of an Australian tarantula. PLoS One 8, e73136.
Hernandez-Campuzano, B., Suarez, R., Lina, L., Hernandez, V., Villegas, E., Corzo, G.,
Iturriaga, G., 2009. Expression of a spider venom peptide in transgenic tobacco confers
insect resistance. Toxicon 53, 122128.
Ikonomopoulou, M., King, G.F., 2013. Natural born insect killers: spider-venom peptides
and their potential for managing arthropod pests. Outlooks Pest Manag. 24, 1619.
Inceoglu, A.B., Kamita, S.G., Hammock, B.D., 2006. Genetically modified baculoviruses:
a historical overview and future outlook. Adv. Virus Res. 68, 323360.
James, C., 2012. Global Status of Commercialized Biotech/GM Crops: 2012. ISAAA
Brief 44-2012. International Service for the Acquisition of Agri-Biotech Applications,
Ithaca, NY.
Jiang, H., Zhu, Y.X., Che, Z.L., 1996. Insect resistance of transformed tobacco plants with the
gene of the spider insecticidal peptide. J. Integr. Plant Biol. (Acta Bot. Sin.) 38, 9599.
410 Volker Herzig et al.
Kennedy, R.M., Steinbaugh, B.A., 2013. Insecticidal triazines and pyrimidines. U.S. Patent
Application 8389718 B2.
Khan, S.A., Zafar, Y., Briddon, R.W., Malik, K.A., Mukhtar, Z., 2006. Spider venom toxin
protects plants from insect attack. Transgenic Res. 15, 349357.
King, G.F., 2011. Venoms as a platform for human drugs: translating toxins into therapeutics.
Expert Opin. Biol. Ther. 11, 14691484.
King, G.F., Hardy, M.C., 2013. Spider-venom peptides: structure, pharmacology, and
potential for control of insect pests. Annu. Rev. Entomol. 58, 475496.
King, G.F., Tedford, H.W., Maggio, F., 2002. Structure and function of insecticidal neuro-
toxins from Australian funnel-web spiders. J. Toxicol. Toxin Rev. 21, 359389.
Klint, J.K., Senff, S., Saez, N.J., Seshadri, R., Lau, H.Y., Bende, N.S., Undheim, E.A.,
Rash, L.D., Mobli, M., King, G.F., 2013. Production of recombinant disulfide-rich
venom peptides for structural and functional analysis via expression in the periplasm
of E. coli. PLoS One 8, e63865.
Kuhn-Nentwig, L., St ocklin, R., Nentwig, W., 2011. Venom composition and strategies in
spiders: is everything possible? Adv. Insect Physiol. 40, 186.
Lomer, C.J., Bateman, R.P., Johnson, D.L., Langewald, J., Thomas, M.B., 2001. Biological
control of locusts and grasshoppers. Annu. Rev. Entomol. 46, 667702.
Maggio, F., Sollod, B.L., Tedford, H.W., Herzig, V., King, G.F., 2010. Spider toxins and
their potential for insect control. In: Gilbert, L.I., Gill, S.S. (Eds.), Insect Pharmacology:
Channels, Receptors, Toxins and Enzymes. Academic Press, London, pp. 101123.
McKay, G.A., Reddy, R., Arhin, F., Belley, A., Lehoux, D., Moeck, G., Sarmiento, I.,
Parr, T.R., Gros, P., Pelletier, J., Far, A.R., 2006. Triaminotriazine DNA helicase inhib-
itors with antibacterial activity. Bioorg. Med. Chem. Lett. 16, 12861290.
Menzler, S., Bikker, J.A., Suman-Chauhan, N., Horwell, D.C., 2000. Design and biological
evaluation of non-peptide analogues of -conotoxin MVIIA. Bioorg. Med. Chem. Lett.
10, 345347.
Michiels, K., Van Damme, E.J., Smagghe, G., 2010. Plant-insect interactions: what can we
learn from plant lectins? Arch. Insect Biochem. Physiol. 73, 193212.
Miller, W.A., Bonning, B.C., 2007. Plant resistance to insect pests mediated by viral proteins.
U.S. Patent No. 7,312,080.
Moar, W.J., Anilkumar, K.J., 2007. Plant science. The power of the pyramid. Science
318, 15611562.
Mukherjee, A.K., Sollod, B.L., Wikel, S.K., King, G.F., 2006. Orally active acaricidal pep-
tide toxins from spider venom. Toxicon 47, 182187.
Nartey, R., Owusu-Dabo, E., Kruppa, T., Baffour-Awuah, S., Annan, A., Oppong, S.,
Becker, N., Obiri-Danso, K., 2013. Use of Bacillus thuringiensis var israelensis as a viable
option in an Integrated Malaria Vector Control Programme in the Kumasi Metropolis,
Ghana. Parasit. Vectors 6, 116.
Nyffeler, M., Sunderland, K.D., 2003. Composition, abundance and pest control potential of
spider communities in agroecosystems: a comparison of European and US studies. Agric.
Ecosyst. Environ. 95, 579612.
Oerke, E.C., 2006. Crop losses to pests. J. Agric. Sci. 144, 3143.
Omar, A., Chatha, K.A., 2012. National Institute for Biotechnology and Genetic Engineer-
ing (NIBGE): genetically modified spider cotton. Asian J. Manag. Cases 9, 3358.
Pal, N., Yamamoto, T., King, G.F., Waine, C., Bonning, B., 2013. Aphicidal efficacy of
scorpion- and spider-derived neurotoxins. Toxicon 70, 114122.
Pallaghy, P.K., Nielsen, K.J., Craik, D.J., Norton, R.S., 1994. A common structural motif
incorporating a cystine knot and a triple-stranded -sheet in toxic and inhibitory poly-
peptides. Protein Sci. 3, 18331839.
Pardo-Lopez, L., Soberon, M., Bravo, A., 2013. Bacillus thuringiensis insecticidal three-
domain Cry toxins: mode of action, insect resistance and consequences for crop protec-
tion. FEMS Microbiol. Rev. 37, 322.
Spider Venom Peptides as Bioinsecticides 411
Platnick, N.I., 2013. The world spider catalog, version 13.5, American Museum of Natural
History Online at http://research.amnh.org/entomology/spiders/catalog/index.html.
Que, Q., Chilton, M.D., De Fontes, C.M., He, C., Nuccio, M., Zhu, T., Wu, Y.,
Chen, J.S., Shi, L., 2010. Trait stacking in transgenic crops: challenges and opportunities.
GM Crops 1, 220229.
Ripka, A.S., Rich, D.H., 1998. Peptidomimetic design. Curr. Opin. Chem. Biol.
2, 441452.
Roberts, D.W., 1973. Means for insect regulation: fungi. Ann. N. Y. Acad. Sci.
217, 7684.
Saez, N.J., Senff, S., Jensen, J.E., Er, S.Y., Herzig, V., Rash, L.D., King, G.F., 2010. Spider-
venom peptides as therapeutics. Toxins 2, 28512871.
Sandhu, S., Sharma, A.K., Beniwal, K., Goel, G., Batra, P., Kumar, A., Jaglan, S.,
Sharma, A.K., Malhotra, S., 2012. Myco-biocontrol of insect pests: factors involved,
mechanism, and regulation. J. Pathog. 2012, 126819.
Shah, P.A., Pell, J.K., 2003. Entomopathogenic fungi as biological control agents. Appl.
Microbiol. Biotechnol. 61, 413423.
Shah, A.D., Ahmed, M., Mukhtar, Z., Khan, S.A., Habib, I., Malik, Z.A., Mansoor, S.,
Saeed, N.A., 2011. Spider toxin (Hvt) gene cloned under phloem specific RSs1 and
RolC promoters provides resistance against American bollworm (Heliothis armigera). Bio-
technol. Lett. 33, 14571463.
Skaer, H.B., Maddrell, S.H., Harrison, J.B., 1987. The permeability properties of septate
junctions in Malpighian tubules of Rhodnius. J. Cell Sci. 88, 251265.
Smith, J.J., Herzig, V., King, G.F., Alewood, P.F., 2013. The insecticidal potential of venom
peptides. Cell. Mol. Life Sci. 70, 36653693.
Soberon, M., Fernandez, L.E., Perez, C., Gill, S.S., Bravo, A., 2007. Mode of action of
mosquitocidal Bacillus thuringiensis toxins. Toxicon 49, 597600.
Tabashnik, B.E., Brevault, T., Carriere, Y., 2013. Insect resistance to Bt crops: lessons from
the first billion acres. Nat. Biotechnol. 31, 510521.
Tedford, H.W., Fletcher, J.I., King, G.F., 2001. Functional significance of the hairpin in
the insecticidal neurotoxin -atracotoxin-Hv1a. J. Biol. Chem. 276, 2656826576.
Tedford, H.W., Gilles, N., Menez, A., Doering, C.J., Zamponi, G.W., King, G.F., 2004a.
Scanning mutagenesis of -atracotoxin-Hv1a reveals a spatially restricted epitope that
confers selective activity against insect calcium channels. J. Biol. Chem.
279, 4413344140.
Tedford, H.W., Sollod, B.L., Maggio, F., King, G.F., 2004b. Australian funnel-web spiders:
master insecticide chemists. Toxicon 43, 601618.
Tedford, H.W., Steinbaugh, B.A., Bao, L., Tait, B.D., Tempczyk-Russell, A., Smith, W.,
Benzon, G.L., Finkenbinder, C.A., Kennedy, R.M., 2013. In silico screening for
compounds that match the pharmacophore of omega-hexatoxin-Hv1a leads to
discovery and optimization of a novel class of insecticides. Pestic. Biochem. Physiol.
106, 124140.
Tice, C.M., 2001. Selecting the right compounds for screening: does Lipinskis Rule of 5 for
pharmaceuticals apply to agrochemicals? Pest Manag. Sci. 57, 316.
Vilcinskas, A., Matha, V., Gotz, P., 1997. Effects of the entomopathogenic fungus
Metarhizium anisopliae and its secondary metabolites on morphology and cytoskeleton
of plasmatocytes isolated from the greater wax moth, Galleria mellonella. J. Insect Physiol.
43, 11491159.
Vollrath, F., Selden, P., 2007. The role of behavior in the evolution of spiders, silks, and
webs. Annu. Rev. Ecol. Evol. Syst. 38, 819846.
Wang, C., St Leger, R.J., 2007. A scorpion neurotoxin increases the potency of a fungal
insecticide. Nat. Biotechnol. 25, 14551456.
Windley, M.J., Herzig, V., Dziemborowicz, S.A., Hardy, M.C., King, G.F.,
Nicholson, G.M., 2012. Spider-venom peptides as bioinsecticides. Toxins 4, 191227.
INDEX
413
414 Index
Crystal toxins (Cry toxins) (Continued ) Aedes aegypti and Anopheles gambiae,
Cry1Aa, 4950, 50f 263264
Cry1Ac protoxin, 318 Drosophila, 262263
Cry34 protein, 6768 DNA screening
definition and classification, 4243 bioassay, 310311
diversity, 4346 cadherin mutations, 311
domain I, 5051 HaCad, 311
domain II, 5153 PCR method, 311
domain III, 5355 r15 cadherin allele, 311312
-endotoxins, 314 DNA shuffling, 231
enterocyte death, 6667 Dolichus biflorus agglutinin (DBA),
genomic sequencing, 49 1517, 16f
4 helix lining, 66 Dorsal anterior rectum (DAR) cells,
insect-resistance trait, 397 1314, 26f
intoxication process, 5556, 55f Drosophila melanogaster
midgut Cry-binding proteins, 5763 Cry1Ac, 6465
models, 6368 dsRNA, 262
oligomer formation, 317 embryo system, 374375
parasporin, 4748 S2 cells, 275
PCR and next-generation sequencing, 49
proteins and molecules, 6263 E
protoxin, 4950 Egg shell structures, 352353
receptors, 316 -Endotoxins, 178179, 187, 314, 379,
ricin domain, 48 397
sequential binding model, 316 Entomopathogens
signalling pathway, 316 Bacillus thuringiensis (Bt), 397
solubilization and proteolytic processing, baculoviruses, 397398
5657 bioinsecticides, 397398
structural domains, 315316 Cry toxins, 397
Culex -endotoxins, 397
and Aedes species, 127129 fungi and parasites (nematodes), 149150
and Anopheles, 93 Green Muscle, 398
C. pipiens, 97t, 102104, 113114, 132t ISVPs, 398399
C. quinquefasciatus Mycotrol, 398
Cry48Aa/Cry49Aa toxin, 105f Entropic spring mechanism, 355356, 356f
growth and mortality, 102104 Environmental effects, GE crops
maltase 1, 114, 132t agronomic properties, 205
Synergism, 6970 cultivation, 203
larvae, 99 gene flow, 205
mosquitoes, 104 grain importation, 205
Cytolytic (Cyt) proteins hazard testing, 203204
description, 184185 monarch butterflies, larvae, 204
mosquitoes and blackflies, toxicity, 187 non-GE isoline, 205
structure and function, 190191, 190f Environmental RNAi
arthropod pests, 269270
D Coleoptera, 253262
Dibrotica virgifera virgifera, 3031, 253 components, 252253
Diptera definitions, 252
Index 417
W Z
Wall-associated protein A (WapA), 352353 Ziconotide/Prialt, 401