2012-Bacterium Lysinibacillus Sphaericus

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Invertebrate Pathology 109 (2012) 1–10

Contents lists available at SciVerse ScienceDirect

Journal of Invertebrate Pathology


journal homepage: www.elsevier.com/locate/jip

Minireview

The bacterium, Lysinibacillus sphaericus, as an insect pathogen


Colin Berry
Cardiff School of Biosciences, Cardiff University, Park Place, Cardiff CF10 3AT, UK

a r t i c l e i n f o a b s t r a c t

Article history: Since the first bacteria with insecticidal activity against mosquito larvae were reported in the 1960s,
Received 16 August 2011 many have been described, with the most potent being isolates of Bacillus thuringiensis or Lysinibacillus
Accepted 12 October 2011 sphaericus (formerly and best known as Bacillus sphaericus). Given environmental concerns over the
Available online 23 November 2011
use of broad spectrum synthetic chemical insecticides and the evolution of resistance to these, industry
placed emphasis on the development of bacteria as alternative control agents. To date, numerous com-
Keywords: mercial formulations of B. thuringiensis subsp. israelensis (Bti) are available in many countries for control
Lysinibacillus sphaericus
of nuisance and vector mosquitoes. Within the past few years, commercial formulations of L. sphaericus
Bacillus sphaericus
Insecticidal toxin
(Ls) have become available. Because Bti has been in use for more than 30 years, its properties are well
Bin toxin know, more so than those of Ls. Thus, the purpose of this review is to summarise the most critical aspects
Cry toxin of Ls and the various proteins that account for its insecticidal properties, especially the mosquitocidal
activity of the most common isolates studied. Data are reviewed for the binary toxin, which accounts
for the activity of sporulated cells, as well as for other toxins produced during vegetative growth, includ-
ing sphaericolysin (active against cockroaches and caterpillars) and the different mosquitocidal Mtx and
Cry toxins. Future studies of these could well lead to novel potent and environmentally compatible insec-
ticidal products for controlling a range of insect pests and vectors of disease.
Ó 2011 Elsevier Inc. All rights reserved.

Contents

1. Introduction and taxonomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.1. Taxonomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Insecticidal activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3. The toxins of L. sphaericus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3.1. Sphaericolysin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
3.2. Vegetative mosquitocidal toxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.2.1. Mtx1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3.2.2. Mtx2 family toxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.3. Bin toxins. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.3.1. Bin structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.3.2. Mechanism of action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.3.3. Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.4. Cry48/Cry49 toxin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
4. Toxin synergy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
5. Field use of L. sphaericus in mosquito control programmes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
6. Genomic analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
7. Future perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1. Introduction and taxonomy known is Bacillus thuringiensis. There are now over 70 subspecies of
B. thuringiensis, with the most well known being those that pro-
Many species of bacterial pathogens have been isolated from in- duce insecticidal endotoxins toxic to the larvae of lepidopteran,
sects over the past century, and by far the most widely studied and coleopteran, or dipteran insects. The subspecies insecticidal for
E-mail address: Berry@cf.ac.uk
dipteran insects, such as B. thuringiensis subsp. israelensis, are

0022-2011/$ - see front matter Ó 2011 Elsevier Inc. All rights reserved.
doi:10.1016/j.jip.2011.11.008
2 C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10

typically most toxic to members of the suborder Nematoceran, the strains tested appearing near-clonal. This confirms previous obser-
so-called long-horned flies, insects such as mosquitoes, blackflies, vations that serotype H5a,5b appears to be a major clone of L. sphae-
and the non-biting chironomid midges. These bacteria have proven ricus amongst the mosquito pathogenic isolates (Zahner et al.,
very efficacious and as a result there are now over one hundred 1998). This serotype contains the strains commercialised for mos-
commercial products on the market for controlling various agricul- quito control such as strains 2362 and C3-41.
tural pests and vectors of human diseases. Moreover, several Cry
proteins, the principal insecticidal components of most isolates of
2. Insecticidal activity
B. thuringiensis, have been engineered into crop plants such as cot-
ton and maize, making the crops tolerant to insect attack.
Since the initial discovery of L. sphaericus insect pathogenicity in
Another insecticidal bacterium that is increasing in importance,
1965 (Kellen et al., 1965), studies have shown mosquitoes to be the
and which has been commercialised over the past decade, is Bacil-
major targets of this bacterium. Reports of activity against other
lus sphaericus, many isolates of which have proven active against
species include lethal effects on eggs of the nematode Trichostron-
the larvae of mosquitoes. It is almost 50 years since Kellen et al.
gus colubriformis (Bone and Tinelli, 1987) and effects on the grass
(1965) first described strains of B. sphaericus as insect pathogens.
shrimp Palaemonetes pugio (Key and Scott, 1992). In addition, indi-
Although this discovery prompted further investigation of the bac-
vidual toxins have been shown to have effects on other species.
terium, it was not until 1973 that strains with the potential for use
Such toxins include the Mtx1 toxin, which has shown activity
in insect control programmes were discovered (Singer, 1973). Since
against Chironomus riparius larvae (Partridge and Berry, 2002)
that time there have been significant advances in our understand-
and sphaericolysin, which is lethal by injection to the German
ing of this bacterium, its taxonomy and the mechanisms by which
cockroach Blattela germanica and the common cutworm Spodoptera
pathogenic strains exert their activity. This review will explore our
litura (Nishiwaki et al., 2007). Subtoxic effects of L. sphaericus,
current understanding of this organism, now renamed Lysinibacil-
including retarded growth and decreased fecundity, have also been
lus sphaericus, with a focus on pathogenic strains and the toxins
reported against the hemipteran water scorpion Laccotrephes gris-
that they produce.
eus, which is a predator of mosquito larvae and some activity at
high doses has been reported for other aquatic organisms
1.1. Taxonomy
(Mathavan et al., 1987). L. sphaericus has no adverse effects on a
range of other organisms (Brown et al., 2004; Lacey, 2007;
Formerly known as Bacillus sphaericus, the species was defined
Mulligan et al., 1978) including beneficial insects such as honey
as having a spherical terminal spore and by its inability to ferment
bees (Davidson et al., 1977) and eukaryotic microorganisms (Tietze
sugars (White and Lotay, 1980). It was reassigned to the new genus
et al., 1993). None of the wider target effects are considered to
Lysinibacillus on the basis of phenotypic characteristics including
compromise the use of L. sphaericus, which has an excellent safety
the fact that members of this new genus differed from other group
record (Lacey, 2007).
2 bacilli in producing peptidoglycans containing lysine and aspar-
The individual toxins produced by insecticidal L. sphaericus
tic acid (Ahmed et al., 2007).
strains and their characteristics will be explored in detail below.
It has long been recognised that within the species L. sphaericus
there is a great deal of variation. Strains of L. sphaericus can be di-
vided into 5 DNA homology groups (I–V) with group II further divis- 3. The toxins of L. sphaericus
ible into subgroups IIA and IIB (Krych et al., 1980). The relatively low
levels of homology between groups led to the suggestion that each Many L. sphaericus strains produce bacteriocins, protein antibi-
might represent a distinct species but as a result of the lack of con- otics that are effective against other strains of the same species
venient tests to distinguish them, all remain designated as L. sphae- (Cokmus and Yousten, 1993) and these could be viewed as antibac-
ricus. Phenotypic analysis does allow some subgroups of L. terial toxins. The focus of the following sections, however, will be
sphaericus to be distinguished (Alexander and Priest, 1990). How- the insecticidal protein toxins produced by insect pathogenic
ever, correlation between phenotype and DNA homology group is strains of this bacterium as described below.
insufficient (Carboulec and Priest, 1989). Nonetheless, phenotypic
traits such as a high likelihood to be resistant to the antibiotics 3.1. Sphaericolysin
chloramphenicol, streptomycin and tetracycline (Alexander and
Priest, 1990), combined with the ability to use arginine as a sole car- Sphaericolysin shows injection toxicity to the German cock-
bon source, have been used to develop a medium to enrich for mos- roach Blattela germanica and the common cutworm Spodoptera lit-
quito pathogenic strains (Yousten et al., 1985). Further phenotypes ura and was discovered in a DNA subgroup IIA L. sphaericus strain
of potential importance for this bacterium include the production by A3-2, lacking bin, mtx1 and mtx2 genes (Nishiwaki et al., 2007).
many strains of restriction endonucleases (Zahner and Priest, 1997) Interestingly, a protein with a high level of sequence identity to
as this could produce a barrier to genetic manipulation. However, sphaericolysin has also been reported in a DNA homology group
strains such as 2297 are naturally restriction negative and the pro- V L. sphaericus strain as a cholesterol-dependent cytolysin (From
duction of restriction deficient derivatives of other strains is also et al., 2008) and sphaericolysin is also encoded in the genome-
possible (Taylor and Burke, 1990). All mosquito pathogenic strains sequenced L. sphaericus strain C3-41 (Hu et al., 2008) so it appears
that have been tested, fall within DNA homology subgroup IIA that the gene encoding this protein may be widespread. Sphaeric-
(Alexander and Priest, 1990; Krych et al., 1980) although not all olysin is secreted into the culture medium, following the cleavage
strains in this group are insect pathogens (Rippere et al., 1997) so of its N-terminal signal sequence, as a 53 kDa protein (Nishiwaki
this feature is not sufficient to distinguish subgroup members. Fur- et al., 2007). The protein is a member of the Thiol cytolysin super-
ther attempts to subdivide strains have employed H-flagellar anti- family of toxins (pfam 01289) that includes perfringolysin O and
gen serotyping (de Barjac et al., 1980) and phage typing (Yousten, alveolysin from B. thuringiensis. It contains a motif seen in the cho-
1984) and there is some correlation between groups assigned by lesterol dependent cytolysins and its insecticidal activity has been
these typing systems and insect pathogenicity. It is interesting to shown to be significantly reduced by co-administration with cho-
note that a recent multi-locus sequence typing study (Ge et al., lesterol. Transmission electron microscopy showed that the toxin
2011) has indicated that there is considerably more heterogeneity was able to produce many pores of around 35 nm diameter in
amongst non-toxic strains than amongst toxin ones, with the toxic erythrocytes, consistent with a perfringolysin O-like mechanism
C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10 3

Fig. 1. Structure of the Mtx1 toxin. Mtx1 structure from PDB accession number 2VSE (Treiber et al., 2008). The N-terminal ADP-ribosyl transferase domain (Red) has the four
ricin-like domains (shades of green and blue) wrapped around it. The loop that is cleaved to activate the toxin and liberate the ADP-ribosyl transferase region is not visible in
the structure, indicating its flexibility and accessibility to its environment. The residues either side of this activation loop are marked with spheres (red and green). The region
C-terminal to the activation region (yellow) binds to the putative NAD+ binding site leading to autoinhibition of activity. The ADP-ribosylating-turn-turn motif that is likely to
determine specificity for the target molecule is also partly missing, indicating flexibility. Those parts visible in the structure are coloured violet and the catalytically important
glutamate residues 195 and 197 are shown in spacefill form.

of action. The A3-2 strain was isolated from the insect Myrmeleon currently unknown. In addition to a low level of toxin synthesis
bore that injects its prey with gut fluid. It is possible, therefore, that in L. sphaericus, Mtx1 is also unstable in this bacterium (Myers
the L. sphaericus strain is able to assist in prey killing by toxin pro- and Yousten, 1978) through degradation by a subtilisin-family ser-
duction and secretion of sphaericolysin by the bacteria growing in ine proteinase named sphaericase (or sfericase) (Wati et al., 1997).
S. litura has been demonstrated in vivo (Nishiwaki et al., 2007). This enzyme (Almog et al., 2003) and its action against Mtx1 has
been the subject of further analysis (Yang et al., 2007) and expres-
3.2. Vegetative mosquitocidal toxins sion of Mtx1 in a DNA homology group IV L. sphaericus strain has
been reported to increase the stability of the protein (Thanabalu
L. sphaericus strains may begin to produce insecticidal toxins and Porter, 1995).
during the vegetative phase of growth. Two classes of vegetative Mtx1 is produced as a 100 kDa protein with an N-terminus that
mosquitocidal toxins (Mtx) have been identified. These can be cat- has features characteristic of a signal sequence although there is no
egorised into two mechanistic groups as follows: evidence that Mtx1 is secreted into the culture medium (Thanab-
alu et al., 1991). Removal of this sequence is required for efficient
3.2.1. Mtx1 expression of the protein in recombinant form and allows produc-
The Mtx1 protein was first identified in the low toxicity strain tion of a 97 kDa toxin in Escherichia coli (Thanabalu et al., 1992a) or
SSII-1 but was found to be widespread in high- and low-toxicity other bacteria (Thanabalu et al., 1992b). The region downstream of
strains (Thanabalu et al., 1991) although in at least one of these the signal peptide-like sequence shows similarities to ADP-ribosyl
strains it appears to be present as a pseudogene (Hu et al., 2008). transferase bacterial toxins (Thanabalu et al., 1991). Proteolytic
Its transcription appears to be active in early exponential growth processing of Mtx1 by enzymes in the mosquito gut or by trypsin
(Ahmed et al., 1995) and the level of production of beta-galactosi- or chymotrypsin, produces a product of 27 kDa containing the
dase, driven by the mtx1 promoter, is lower in L. sphaericus than in ADP-ribosyl transferase sequences and a product of 70 kDa with
Bacillus subtilis, suggesting the action of a regulatory element in the internal repeat sequences (Schirmer et al., 2002a; Thanabalu et al.,
natural host (Ahmed et al., 1995). An inverted repeat sequence up- 1992a). These repeat sequences represent lectin-like motifs (Hazes
stream of the initiation codon of the mtx1 gene was initially and Read, 1995) that may have a role in binding to sugar groups
thought to be a regulator binding region (Thanabalu et al., 1991). and, thus, in determining the specificity of the toxin. The crystal
Analysis of the region immediately upstream of the mtx1 gene in structure of the toxin has been elucidated and shows four ri-
the L. sphaericus C3-41 genome (Hu et al., 2008) reveals a sensory cin B-like lectin domains surrounding the ADP-ribosyl transferase
box protein (often involved in detection of external conditions domain in a way that is likely to cause allosteric inhibition of its
and signaling), preceded by a BglG anti-terminator family protein. enzyme activity (Treiber et al., 2008) (Fig. 1). The lectin domains
Inspection of the inverted repeat upstream of mtx1 and regulatory are related to pierisin cytotoxin, which is known to bind glycolip-
inverted repeat sequences in the bgl operon of E. coli (Houman ids and the authors proposed that Mtx1 might utilise a similar
et al., 1990) reveal similarities in location and length. In the bgl op- receptor. The activation loop in the crystal structure is exposed
eron there is a similar arrangement of genes with the BglG anti- and would allow proteolytic cleavage to allow the lectin domain
terminator encoded upstream of the BglF regulator, followed by to mediate uptake of the catalytic domain into the target cell in a
the gene that, together, they regulate. Further similarity lies in manner typical of this class of toxin. Although the target(s) of
the fact that in the bgl operon there is a second stem loop between Mtx1 in mosquito cells are unknown, Mtx1 modifies unknown tar-
the anti-terminator gene and the regulator gene and an inverted get proteins in Culex quinquefasciatus (Thanabalu et al., 1993), sev-
repeat sequence is also present in the mtx region between the eral proteins in HeLa cells, EF-Tu in E. coli (Schirmer et al., 2002b)
two upstream operons. Hence, it is possible that mtx1 gene expres- and soybean trypsin inhibitor, apparently targeting arginine
sion is regulated in a similar manner involving anti-termination residues for modification (Carpusca et al., 2006; Schirmer et al.,
but the nature of the signal to which this system may respond is 2002a). However, the kinetics of this reaction with respect to
4 C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10

Fig. 2. Alignment of Mtx2 family sequences. Sequences aligned using ClustalX. Shaded residues indicate identity between sequences.

NAD+ (KM = 45 lM; kcat = 0.04 s 1) (Schirmer et al., 2002a) show a activity (Phannachet et al., 2010). To date, the activity of the
somewhat greater affinity but a far lower turnover for this sub- Mtx4 putative toxin has not been assessed.
strate under the experimental conditions than those reported for
the related ADP-ribosyltransferase pierisin (KM = 170 lM; 3.3. Bin toxins
kcat = 55 s 1) (Watanabe et al., 2004).
The C-terminal 70 kDa region of Mtx1 (containing the lectin re- The Bin or Binary toxin is deposited in the form of a parasporal
peats) is able to cause morphological changes in both C. quinquefas- crystal and is found characteristically in high-toxicity mosquitoci-
ciatus and Aedes aegypti cells in culture but both the 27 kDa dal strains (Fig. 3). The toxin comprises two proteins, BinA (370
enzymatic domain and the 70 kDa domain are necessary for toxic- amino acids, 42 kDa) and BinB (448 amino acids, 51 kDa)
ity to mosquito larvae (Thanabalu et al., 1993). The ADP-ribosylat- (Baumann et al., 1988; Hindley and Berry, 1987) that are co-tran-
ing 27 kDa domain is able to cause cytotoxicity when expressed scribed from a single operon just before the end of exponential
inside HeLa cells and cells developed actin-containing structures growth and into sporulation (Ahmed et al., 1995; Baumann and
resembling filopodia and rounding of cells (Schirmer et al., 2002a). Baumann, 1989) and strains blocked in early stage sporulation do
not accumulate toxin crystals (Charles et al., 1988; El-Bendary
et al., 2005). The Bin protein sequences are highly conserved
3.2.2. Mtx2 family toxins between strains with only five variants reported, differing by no
The Mtx2 protein was also originally identified in L. sphaericus more than six amino acids in each protein between any two vari-
strain SSII-1 (Thanabalu and Porter, 1996) and is unrelated to the ants (Hire et al., 2009; Humphreys and Berry, 1998; Priest et al.,
Mtx1 protein. Subsequently, a further member of this family,
Mtx3, was characterised (Liu et al., 1996a) and sequencing of the
L. sphaericus C3-41 genome showed that it encoded a further re-
lated protein, Mtx4, along with an apparent pseudogene member
of this family (Hu et al., 2008) (Fig. 2). Thus, it appears that there
has been considerable duplication within this gene family, which
is represented in many high- and low-toxicity L. sphaericus strains.
The proteins are members of the family of pore forming toxins de-
fined in the NCBI Conserved Domain Database (Marchler-Bauer
et al., 2011) as the Clostridium epsilon toxin ETX/MTX2 family
(pfam 03318). Related proteins including Cry15Aa are also pro-
duced by B. thuringiensis strains (de Maagd et al., 2003). Mtx2
and Mtx3 appear to be produced during vegetative growth and
exhibit mosquitocidal activity against C. quinquefasciatus and
A. aegypti (Liu et al., 1996a; Thanabalu and Porter, 1996). Activity
of the Mtx2 toxin towards the two mosquito species is different
for natural variants of the toxin and amino acid residue 224 has
been shown to be critical in determining the optimal target: lysine
Fig. 3. Electron micrograph of sporulated Lysinibacillus sphaericus cells. The small
at this position favours activity against C. quinquefasciatus whereas dark inclusions (arrows), which adhere to the internal surface of the exosporium
threonine favours activity against A. aegypti (Chan et al., 1996). The membrane, are inclusions consisting of BinA and BinB. (Image kindly provided by
C-terminus of Mtx2 may also be important for its solubility and Dr. J.-F. Charles.)
C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10 5

1997). The two proteins share regions of identity and appear to caecum and posterior midgut in Culex mosquitoes, less regionally
have evolved from a common ancestor (Baumann et al., 1988). in Anopheles and not at all in Aedes aegypti (Davidson, 1988,
Other members of the Bin toxin family (pfam 05431) include the 1989; Oei et al., 1992). The binding in Culex is mediated by the BinB
Cry49 protein from L. sphaericus (see below); the B. thuringiensis protein, which can bind alone to the distinct regions of the gut (Oei
toxins Cry35 and Cry36; and a putative protein from Chlorobium et al., 1992) and this protein associates specifically with a single
phaeobacteriodes (Jones et al., 2007). receptor (Charles et al., 1997) subsequently identified as a GPI-
On ingestion by larvae, the Bin toxin crystal is solubilised in the anchored maltase, Cpm1 (Culex pipiens maltase 1) (Darboux et al.,
alkaline environment of the gut and the proteins undergo limited 2001; Silva-Filha et al., 1999). Toxin binding to brush border mem-
proteolysis to lower molecular weight derivatives (Aly et al., brane fractions from Anopheles larvae has also been shown
1989; Broadwell and Baumann, 1987; Davidson et al., 1987a; (Silva-Filha et al., 1997) but in Anopheles gambiae, BinA appears
Davidson et al., 1990) with an increased toxicity (Broadwell and to be able to bind independently of BinB although the binding of
Baumann, 1987) and this proteolysis occurs in both susceptible both components seems to be cooperative with maximum binding
and non-susceptible insects (Nicolas et al., 1990). Consistent with observed at a 1:1 molar ratio of BinA and BinB (Charles et al.,
the processing of the toxin, deletion experiments have shown that 1997). The receptor in this species appears also to be a maltase
17 amino acids can be removed from both the N- and C-terminus ortholog of the Cpm1 protein and has been named Agm1 (Opota
of BinA and that 34 residues can be removed from the N-terminus et al., 2008). A further ortholog from Aedes aegypti, a mosquito
and 53 from the C-terminus of BinB, without abolishing toxicity showing little or no sensitivity to Bin toxins, shows a high degree
(Broadwell et al., 1990b; Clark and Baumann, 1990; Oei et al., of conservation with Cpm1 and Agm1, indicating that specificity
1990). may be determined by the high affinity recognition of a relatively
small region that is less highly conserved between species (Opota
3.3.1. Bin structure et al., 2008). Expression of the Cpm1 receptor in mammalian cells
Secondary structure analysis of BinA from its circular dichroism showed that it was polarised to the apical surface and was associ-
spectrum suggests a composition of 49.3% beta strand and 3.1% al- ated with lipid rafts (Pauchet et al., 2005).
pha helix (Hire et al., 2009) whereas BinB is reported to contain a The sequence of events following receptor binding that lead to
high percentage of alpha helix (Tangsongcharoen et al., 2011). cell death are less clear. Patch clamping experiments on Culex quin-
Interaction of BinA and BinB in solution is reported to increase beta quefasciatus cells showed pore formation by the toxin components
sheet composition and further structural changes occur on interac- (Cokmus et al., 1997). Studies using artificial membranes have
tion with lipid membranes (Boonserm et al., 2006). Oligomeri- shown similar results although studies differ with regard to which
sation of BinA and BinB occurs in solution to give a BinA2BinB2 component may be most important for this activity (Boonserm
tetramer (Smith et al., 2005). Some progress towards elucidating et al., 2006; Schwartz et al., 2001). It is, therefore, possible that
a 3-dimensional structure for the Bin proteins has been made with the mode of action of the Bin toxin derives from its pore forming
the crystallisation of the BinB protein alone (Chiou et al., 1999) and ability. Within the intoxicated cells, mitochondrial cristae along
a BinA/BinB co-crystal (Smith et al., 2004) but, to date, no crystal with the endoplasmic reticulum are seen to dilate while the mito-
structure is available. The availability of such a structure will aid chondria themselves condense, there is a marked reduction in
in understanding receptor binding, mechanism of action and the intracellular ATP and a gradual loss of integrity in the cell mem-
structural importance of residues shown to influence toxicity and brane (Broadwell and Baumann, 1986; Charles, 1987; Davidson
target range of the Bin toxin. and Titus, 1987). However, intoxication by Bin shows other inter-
Several amino acids have been identified with importance in esting features: internalisation of the toxin and vacuolation of
maintaining the activity of the Bin proteins (Boonyos et al., 2009; the target cell (Davidson, 1989; Davidson et al., 1987b; Davidson
Elangovan et al., 2000; Promdonkoy et al., 2008; Sanitt et al., and Titus, 1987; Oei et al., 1992). Cloning of the Cpm1 receptor
2008; Shanmugavelu et al., 1998; Yuan et al., 2001) and determin- gene into mammalian cells leads to the opening of pores and the
ing its level of activity against A. aegypti (Berry et al., 1993). Resi- appearance of vacuoles on addition of Bin toxin although cell lysis
due 150 was suggested to be important in receptor binding by does not occur (Pauchet et al., 2005). A detailed study of the action
BinB (Singkhamanan et al., 2010) and further, recent work has of Bin toxin on these cells showed the induction of cationic pores,
identified the region from residues 33–158 to be important for internalisation of both toxin components along with their receptor
receptor binding, with residues 147–149 being critical to binding and the production of large vacuoles (Opota et al., 2011). These
(Romao et al., 2011). In another study, mutants in the N- and C-ter- vacuoles, however, are not the location of the internalised toxins.
mini of each protein that alone were found to eliminate toxicity The toxin-receptor complex is endocytosed and trafficked into
were found to complement each other to restore toxicity when recycling endosomes, whereas the distinct vacuoles are autophagic
mixed (Shanmugavelu et al., 1998). This was taken to suggest olig- in character. One hypothesis is that pore formation may be sensed
omerisation of the toxin subunits and this phenomenon was subse- by p38 mitogen-activated protein kinase, the activity of which
quently shown experimentally (Smith et al., 2005) (see above). might induce autophagy. The vacuoles are transient but were ob-
served to reform in cells following cell division, a phenomenon that
3.3.2. Mechanism of action was termed post-mitotic vacuolation. Bin toxin thus causes vacuo-
The Bin toxin is generally considered a binary toxin since both lation but its own trafficking is distinct and avoids targeting to deg-
components are necessary for full toxicity (Broadwell et al., radative organelles. The significance for the toxicity of Bin proteins
1990a; Oei et al., 1990) with optimum potency produced at a of each of the individual responses it elicits, at present, remains un-
1:1 M ratio of the two proteins. However, activity of the BinA pro- clear. It has been speculated that the trafficking of Bin toxin may
tein alone expressed in B. thuringiensis has been observed allow it to gain access to deeper tissues (Opota et al., 2011), a factor
(Baumann and Baumann, 1991; Nicolas et al., 1993). The primary that could explain muscle and nerve damage following larval expo-
site of action of the toxin is the insect gut after ingestion by filter sure to the toxins (Singh and Gill, 1988).
feeding larvae but intoxicated larvae also show damage to nervous
and muscle tissues (Singh and Gill, 1988). 3.3.3. Resistance
The fact that most L. sphaericus strains produce only one spore
3.3.2.1. Receptor binding. Following solubilisation and activation in associated toxin (Bin) and that spores are the agent utilised in mos-
the mosquito gut, toxin binds to epithelial cells in the gastric quito control programmes, means that it may be relatively easy for
6 C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10

insects to evolve high-level resistance (Nielsen-Leroux et al., 1995; resistance and screening for strains with higher-level production
Wirth et al., 2000b). However, reductions in fecundity and fertility of this new toxin may be worthwhile.
may arise (de Oliveira et al., 2003; Rodcharoen and Mulla, 1997)
and resistance is recessive (Oliveira et al., 2004; Wirth et al.,
4. Toxin synergy
2000b) so strategies for controlling resistance in the field are viable
(Mulla et al., 2003).
Both the vegetative toxins Mtx1 and Mtx2 are able to act in syn-
Apocrine secretions by mosquito gut cells have been proposed
ergism against Aedes aegypti (Rungrod et al., 2009) and each acts
as possible mechanisms by which the insects may seek to defend
synergistically with the Bin toxin (Wirth et al., 2007) and may be
themselves against L. sphaericus (Oliveira et al., 2009). However,
valuable in enhancing the effectiveness of L. sphaericus spore prep-
most cases of resistance to this bacterium and the Bin toxin that
arations. Although the Mtx proteins are produced during vegeta-
have been reported to date, involve changes to its receptor in the
tive growth and Bin proteins are associated with the spores, it
gut, which abolish toxin localisation to host cell membranes. The
has been shown that, in intoxicated insects, most L. sphaericus have
mechanisms of resistance include a variety of mis-sense mutations
entered vegetative growth by the time that the larvae die (Charles,
that remove the GPI anchor sequence from the receptor and inser-
1987). Thus, it is possible that the production of vegetative toxins
tion of a transposable element into the receptor gene to cause al-
and their synergism with Bin may be physiologically relevant in
tered splicing leading to the production of a protein that contains
the natural progression of a L. sphaericus infection. Although
the GPI anchor sequence but carries an internal deletion of 66 ami-
Mtx1, Mtx2 and Mtx3 are associated with the vegetative cells
no acids (Darboux et al., 2002, 2007; Romão et al., 2006). A further
and do not appear to be secreted, early experiments with strain
resistant mosquito collected from field treatment areas in France
SSII-1, which does not produce Bin toxins, suggested that digestion
exhibited high level resistance (>10,000 fold) but with no loss of
of the bacteria in the gut might serve to release toxins (Davidson,
receptor binding (Nielsen-Leroux et al., 1997). Strains resistant to
1979). Synergistic actions of L. sphaericus toxins and B. thuringiensis
one Bin variant are cross-resistant to all other variants (Nielsen-
insecticidal toxins have also been reported (Mtx1 with Cry11A;
LeRoux et al., 2001; Yuan et al., 2003) and this renders most strains
Mtx2 with Cry11A and Cyt1A (Wirth et al., 2007): Bin with Cry4A,
of L. sphaericus ineffective against these resistant Culex larvae.
Cry4B and Cry11A (Wirth et al., 2004): Bin with Cyt proteins
Some strains, however, remain able to kill Bin resistant larvae
(Wirth et al., 2001, 2000a): and possible interactions between
(Nielsen-LeRoux et al., 2001; Pei et al., 2002; Yuan et al., 2003)
Cry49 and Cry4A (Jones et al., 2008)). The distinct mechanisms of
and this led to the discovery of a further toxin in some L. sphaericus
action of B. thuringiensis toxins and those of L. sphaericus has led
strains.
several groups to experiment with combinations of toxins, ex-
pressed in both bacteria (Gammon et al., 2006; Park et al., 2003;
3.4. Cry48/Cry49 toxin
Poncet et al., 1994, 1997; Yang et al., 2007) and in other microor-
ganisms (Liu et al., 1996b; Tanapongpipat et al., 2003; Tandeau
Those strains able to overcome Bin resistance or against which
de Marsac et al., 1987; Thanabalu et al., 1992b; Xudong et al.,
resistance is slower to develop, produce extra crystal proteins on
1993; Yap et al., 1994). None of these recombinants are currently
sporulation (Pei et al., 2002; Yuan et al., 2003). Analysis of these
used in the field.
strains led to the discovery of a new, two component toxin (Jones
et al., 2007). One component is the Cry48 protein, which is closely
related to the mosquitocidal Cry toxins of B. thuringiensis, and is 5. Field use of L. sphaericus in mosquito control programmes
clearly a member of the 3-domain Cry toxin family. The second
is the Cry49 protein, which is a member of the Bin family toxins L. sphaericus preparations are produced commercially for use in
(see above). Neither component shows toxicity alone but, in com- vector control programmes (e.g. GriselESF from Labiofam, Cuba;
bination, the purified proteins exhibit high-level toxicity to Culex Sphaerus from Bthek Ltda, Brazil; VectoLex and VectoMax -a mix-
larvae although their known target range is highly restricted to this ture of L. sphaericus with B. thuringiensis, from Valent Biosciences
genus of insects (Jones et al., 2008). This co-dependence of a Bin- Corporation, USA). The bacterium is applied as a spore preparation
like and a 3-domain Cry toxin is novel (although synergy between that also contains the associated Bin toxin crystals. It is recom-
Bin and B. thuringiensis Cry toxins has been reported (Wirth et al., mended for use against Culex and Anopheles mosquitoes but while
2004)). Analysis of the effects of the Cry48/Cry49 toxin on Culex it has little or no activity against Aedes aegypti, it provides good
larvae showed morphological changes in target cells that mim- control for Aedes vexans, Ochlerotatus atropalpus, Ochlerotatus fitch-
icked the synergistic interaction of Bin and Cry11 toxins with some ii, Ochlerotatus intrudens, Ochlerotatus nigromaculis and Ochlerotatus
changes typical of Bin intoxication (e.g. cytoplasmic vacuolization, stimulans (Mulligan et al., 1978; Wraight et al., 1987). Ochlerotatus
fragmentation of endoplasmic reticulum) and some typical of has been proposed both as a relatively new genus, members of
3-domain Cry toxin activity (e.g. mitochondrial swelling) (de Melo which were formerly included in the genus Aedes (Reinert, 2000)
et al., 2009). Thus, the combination appears to show aspects to its and as a subgenus of Aedes (Savage and Strickman, 2004). Thus,
toxicity with features of both toxin classes. To date, the roles of the generality that L. sphaericus is not a good control agent for
each component in the processes of receptor binding and toxicity Aedes mosquitoes does not hold true for all Aedes species, particu-
are still to be determined. larly those in the genus/subgenus Ochlerotatus (although there are
The toxicity of the combined proteins to Culex larvae is high but, conflicting reports concerning activity against Ochlerotatus taenio-
in the host L. sphaericus strains, toxicity is limited by the low-level rhynchus (Hayes et al., 2011; Ramoska et al., 1977)). The bacteria
production of the Cry48 component. The arrangement of the cry48 also exhibit toxicity to Mansonia and Psorophora mosquitoes
and cry49 genes as convergent transcripts, along with the low lev- (Cheong and Yap, 1985; Ramoska et al., 1977) but have little or
els of production has led to speculation that this combination may no activity against most Toxorhynchites species, the larvae of which
be a relatively recent event (Jones et al., 2007). Mosquitoes with may be predators of other mosquito larvae (Lacey, 1983; Lacey
resistance to strains producing both Bin and Cry48/Cry49 (such et al., 1988; Thanabalu et al., 1992a).
as strain IAB59) can be produced by selection but require longer- In addition to the well-known toxicity of L. sphaericus to mos-
term pressure and resistance arises more slowly (Amorim et al., quito larvae, a reduction in oviposition and the death of adult Culex
2007). As a result, to control Culex larvae, strains producing both quinquefasciatus mosquitoes has also been reported (Zahiri and
Bin and Cry48/Cry49 may be valuable to reduce problems of Mulla, 2005) and this may be an important effect in helping to
C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10 7

reduce disease transmission. The mechanism of adult killing has (Han et al., 2007; Hu et al., 2008). Some toxin genes are closely lo-
not been determined but adult sensitivity in some circumstances cated with respect to others (bin with mtx4; mtx1 with mtx2; mtx3
has been reported although L. sphaericus did not appear orally ac- with the mtx2/3-like pseudogene) but there is no single ‘‘pathoge-
tive in adults (Stray et al., 1988). Further effects on adult mosqui- nicity island’’. Toxin genes frequently have nearby transposase
toes may include a decrease in the potential to transmit malarial genes and, in the case of the bin/mtx4x genes, nearby phage integr-
parasites as shown with the rodent malaria parasite Plasmodium ase-like sequences, perhaps indicating mechanisms by which toxin
berghei and the mosquito Anopheles quadrimaculatus following lar- genes may have been acquired and explaining the distribution of
val exposure to the spores (Young et al., 1990). genes in different strains. The mtx1 gene in this strain, in contrast
Advantages of L. sphaericus in control programmes include gen- to the SSII-1 strain, appears to be a pseudogene. In addition, the po-
erally good persistence, particularly in polluted waters, which are tential control region upstream of the mtx1 gene (see above) is al-
good breeding sites for Culex mosquitoes (Des Rochers and Garcia, tered in strain C3-41 so that it no longer represents an exact
1984; Karch et al., 1988; Nicolas et al., 1987) and this persistence is inverted repeat. This may also be a sign that the mtx1 gene in this
higher than that of B. thuringiensis subsp. israelensis (Sil- strain is no longer active.
apanuntakul et al., 1983). There was no significant difference be- Strain C3-41 also contains a major extrachromosomal element,
tween the stability of Bin toxins in varying temperatures and the 177 kb pBsph plasmid. This plasmid contains a 30.5 kb duplica-
water quality when produced in L. sphaericus or in recombinant tion of the genome and this region includes the bin and mtx4 genes.
B. thuringiensis where they are deposited inside or outside of the Duplication of the bin genes in strain 2297 has also been suggested
exosporium respectively. This indicates that the exosporium does previously (Poncet et al., 1997). The pBsph plasmid was the only
not play a role in stabilizing the toxin (Nicolas et al., 1994). How- extrachromosomal element characterised during the genome
ever, toxins produced in either bacterium showed decreased stabil- sequencing of strain C3-41 but L. sphaericus strains may contain a
ity as temperature increased and water quality declined. range of plasmids of unknown function (Abe et al., 1983; Wu
Persistence may be related to UV resistance and sensitivity to et al., 2007). Since the Cy48/Cry49 toxin is not produced by strain
UV-B for both the spore and the toxin have been reported (Hada- C3-41, at present, the location of the cry48/cry49 genes in L. sphae-
pad et al., 2008). However, insecticidal activity shows persistence ricus strains has not been determined.
even after spore viability has declined (Burke et al., 1983) and sta-
bility of both spore and crystal components can be enhanced to
some extent by the addition of UV protectants (Hadapad et al.,
2009). L. sphaericus spores also show a lower tendency than B. thur- 7. Future perspectives
ingiensis spores to stick to and sediment with other particulates in
the water column, perhaps increasing their persistence in the lar- L. sphaericus has proved to be a highly effective mosquito con-
val feeding zones (Yousten et al., 1992). However, sedimentation trol agent in the field despite the fact that resistance can occur.
of spores does generally remove them from feeding zones (David- Management of this potential problem, for instance by use in com-
son et al., 1984) although larvae grazing on shallow muds may still bination or rotation with other agents such as B. thuringiensis, is
be exposed (Matanmi et al., 1990) and this may be important for likely to maintain the utility of L. sphaericus pesticides. Our under-
persistence. standing of this organism and its insect pathogenicity has in-
There is also evidence that L. sphaericus recycles in the environ- creased greatly in recent years but a number of issues remain to
ment. The bacteria do not recycle in the water in treated ponds and be clarified. While we are able to use this bacterium in our control
there is no evidence of persistence in pools that dry out and are programmes by the application of high densities of spores in the
subsequently re-flooded (Davidson et al., 1984). Spores ingested field, much less is known of the natural ecology of this organism.
by invertebrates other than mosquitoes remain viable although Its persistence and recycling may be important under natural con-
they do not appear to reproduce. There is also evidence that mid- ditions to allow the lower doses of bacteria normally available in
ges emerging from treated water might serve to carry spores to the environment to initiate infections in a population of insects.
new locations (Yousten et al., 1991). Bacterial growth does occur This may be an interesting subject for further study.
in mosquito larval cadavers and bacteria are released when these At the molecular level, although important information on the
breakup (Becker et al., 1995; Des Rochers and Garcia, 1984). The mechanism of action of the toxins has been elucidated, a detailed
recycling ability seems to be a feature of high toxicity, Bin produc- understanding is still elusive. The evidence suggests that the Bin
ing strains that are able to germinate rapidly in the insect and the toxin can form pores and we have begun to understand the process
final L. sphaericus cell counts in cadavers have been estimated in by which vacuolization of the target cell occurs but the significance
the range of approximately 1–1000 times the number of spores of these phenomena in the actual progress of toxicity must be clar-
originally ingested (Charles and Nicolas, 1986; Correa and Yousten, ified. In the Cry48/Cry49 system, the role played by each compo-
1995; Davidson et al., 1984). Spores produced in the dead insect nent in receptor binding and toxicity should be investigated
are associated with Bin crystals and show the same level of toxicity further. Progress in the structural analyses of the toxins may allow
as cultures grown in artificial media. This production in larvae may a better understanding of the receptor specificities and mecha-
contribute to the persistence of the control in the field but only if nisms of action. The 3-dimensional structure of Mtx1 is well stud-
the cadavers remain in the larval feeding zone in the upper water ied but for the other toxins, there is still much to be discovered.
layers (Charles and Nicolas, 1986) or in shallow waters (Kramer, Receptors for Mtx1 and Mtx2-family proteins have yet to be iden-
1990). tified and the molecular target for ADP-ribosylation by Mtx1 in
mosquito cells requires further investigation.
The future development of recombinant L. sphaericus or other
6. Genomic analysis organisms carrying its toxins may allow production of strains with
higher potency, that are better able to delay resistance and/or with
Analysis of the genome of the high toxicity L. sphaericus strain longer persistence in the field. Such developments will be driven
C3-41 reveals a sequence of 4.6 Mb (Hu et al., 2008). The bacterium by our increasing understanding of the bacterium and its toxins
lacks genes encoding sugar transporters (with the exception of but will also be mediated by the attitudes of the regulatory bodies
N-acetyl glucosamine) and proteins involved in sugar metabolism, and the willingness of the public at large to accept field use of re-
explaining the inability of L. sphaericus to grow on many sugars combinant organisms. Whatever form it may take, the continued
8 C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10

use of L. sphaericus in mosquito control programmes seems assured Cheong, W.C., Yap, H.H., 1985. Bioassays of Bacillus sphaericus (strain 1593) against
mosquitoes of public health importance in Malaysia. Southeast Asian J. Trop.
for the foreseeable future.
Med. Public Health 16, 54–58.
Chiou, C., Davidson, E.W., Thanabalu, T., Porter, A.G., Allen, J.P., 1999. Crystallization
References 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, 1083–1085.
Abe, K., Faust, R.M., Bulla, L.A., 1983. Plasmid deoxyribonucleic acid in strains of
Clark, M.A., Baumann, P., 1990. Deletion analysis of the 51-kilodalton protein of the
Bacillus sphaericus and in Bacillus moritai. J. Invertebr. Pathol. 41, 328–335.
Bacillus sphaericus 2362 binary mosquitocidal toxin: construction of derivatives
Ahmed, H.K., Mitchell, W.J., Priest, F.G., 1995. Regulation of mosquitocidal toxin
equivalent to the larva-processed toxin. J. Bacteriol. 172, 6759–6763.
synthesis in Bacillus sphaericus. Appl. Environ. Microbiol. 43, 310–314.
Cokmus, C., Yousten, A.A., 1993. Bacteriocin production by Bacillus sphaericus. J.
Ahmed, I., Yokota, A., Yamazoe, A., Fujiwara, T., 2007. Proposal of Lysinibacillus
Invertebr. Pathol. 61, 323–325.
boronitolerans gen. nov. sp. nov., and transfer of Bacillus fusiformis to
Cokmus, C., Davidson, E.W., Cooper, K., 1997. Electrophysiological effects of Bacillus
Lysinibacillus fusiformis comb. nov. and Bacillus sphaericus to Lysinibacillus
sphaericus binary toxin on cultured mosquito cells. J. Invertebr. Pathol. 69, 197–
sphaericus comb. nov. Int. J. Syst. Evol. Microbiol. 57, 1117–1125.
204.
Alexander, B., Priest, F.G., 1990. Numerical classification and identification of
Correa, M., Yousten, A.A., 1995. Bacillus sphaericus spore germination and recycling
Bacillus sphaericus including some strains pathogenic for mosquito larvae. J.
in mosquito larval cadavers. J. Invertebr. pathol. 66, 76–81.
Gen. Microbiol. 136, 367–376.
Darboux, I., Nielsen-LeRoux, C., Charles, J.F., Pauron, D., 2001. The receptor of
Almog, O., Gonzalez, A., Klein, D., Greenblatt, H.M., Braun, S., Shoham, G., 2003. The
Bacillus sphaericus binary toxin in Culex pipiens (Diptera: Culicidae) midgut:
0.93A crystal structure of sphericase: a calcium-loaded serine protease from
molecular cloning and expression. Insect Biochem. Mol. Biol. 31, 981–990.
Bacillus sphaericus. J. Mol. Biol. 332, 1071–1082.
Darboux, I., Pauchet, Y., Castella, C., Silva-Filha, M. H., Nielsen-LeRoux, C., Charles, J.
Aly, C., Mulla, M.S., Federici, B.A., 1989. Ingestion, dissolution, and proteolysis of the
F., Pauron, D., 2002. Loss of the membrane anchor of the target receptor is a
Bacillus sphaericus toxin by mosquito larvae. J. Invertebr. Pathol. 53, 12–20.
mechanism of bioinsecticide resistance. Proc. Natl. Acad. Sci. USA 99, 5830–
Amorim, L.B., Fontes de Oliveira, C.M., Rios, E.M., Regis, L., Silva-Filha, M.H., 2007.
5835.
Development of Culex quinquefasciatus resistance to Bacillus sphaericus strain
Darboux, I., Charles, J.F., Pauchet, Y., Warot, S., Pauron, D., 2007. Transposon-
IAB59 needs long term selection pressure. Biol. Control. 42, 155–160.
mediated resistance to Bacillus sphaericus in a field-evolved population of Culex
Baumann, L., Baumann, P., 1989. Expression in Bacillus subtilis of the 51- and 42-
pipiens (Diptera: Culicidae). Cell Microbiol. 9, 2022–2029.
kilodalton mosquitocidal toxin genes of Bacillus sphaericus. Appl. Environ.
Davidson, E.W., 1979. Ultrastructure of midgut events in the pathogenesis of
Microbiol. 55, 252–253.
Bacillus sphaericus strain SSII-1 infections of Culex pipiens quinquefasciatus
Baumann, L., Baumann, P., 1991. Effects of components of the Bacillus sphaericus
larvae. Can. J. Microbiol. 25, 178–184.
toxin on mosquito larvae and mosquito-derived tissue culture-grown cells.
Davidson, E.W., 1988. Binding of the Bacillus sphaericus (Eubacteriales: Bacillaceae)
Curr. Microbiol. 23, 51–57.
toxin to midgut cells of mosquito (Diptera: Culicidae) larvae: relationship to
Baumann, L., Broadwell, A.H., Baumann, P., 1988. Sequence analysis of the
host range. J. Med. Entomol. 25, 151–157.
mosquitocidal toxin genes encoding 51.4- and 41.9-kilodalton proteins from
Davidson, E.W., 1989. Variation in binding of Bacillus sphaericus toxin and wheat
Bacillus sphaericus 2362 and 2297. J. Bacteriol. 170, 2045–2050.
germ agglutinin to larval midgut cells of six species of mosquitoes. J. Invertebr.
Becker, N., Zgomba, M., Petric, D., Beck, M., Ludwig, M., 1995. Role of larval cadavers
Pathol. 53, 251–259.
in recycling processes of Bacillus sphaericus. J. Am. Mosquito Control Assoc. 11,
Davidson, E.W., Titus, M., 1987. Ultrastructural effects of the Bacillus sphaericus
329–334.
mosquito larvicidal toxin on cultured mosquito cells. J. Invertebr. Pathol. 50,
Berry, C., Hindley, J., Ehrhardt, A.F., Grounds, T., de Souza, I., Davidson, E.W., 1993.
213–220.
Genetic determinants of the host range of the Bacillus sphaericus mosquito
Davidson, E.W., Morton, H.L., Moffett, J.O., Singer, S., 1977. Effect of Bacillus
larvicidal toxins. J. Bacteriol. 175, 510–518.
sphaericus Strain SSII-1 on Honey Bees, Apis-Mellifera. J. Invertebr. Pathol. 29,
Bone, L.W., Tinelli, R., 1987. Trichostrongylus colubriformis: larvicidal activity of toxic
344–346.
extracts from Bacillus sphaericus (strain 1593) spores. Exp. Parasitol. 64, 514–
Davidson, E.W., Urbina, M., Payne, J., Mulla, M.S., Darwazeh, H., Dulmage, H.T.,
516.
Correa, J.A., 1984. Fate of Bacillus sphaericus 1593 and 2362 spores used as
Boonserm, P., Moonsom, S., Boonchoy, C., Promdonkoy, B., Parthasarathy, K., Torres,
larvicides in the aquatic environment. Appl. Environ. Microbiol. 47, 125–129.
J., 2006. Association of the components of the binary toxin from Bacillus
Davidson, E.W., Bieber, A.L., Meyer, M., Shellabarger, C., 1987a. Enzymatic activation
sphaericus in solution and with model lipid bilayers. Biochem. Biophys. Res.
of the Bacillus sphaericus mosquito larvicidal toxin. J. Invertebr. Pathol. 50, 40–
Commun. 342, 1273–1278.
44.
Boonyos, P., Soonsanga, S., Boonserm, P., Promdonkoy, B., 2009. Role of cysteine at
Davidson, E.W., Shellabarger, C., Meyer, M., Bieber, A.L., 1987b. Binding of the
positions 67, 161 and 241 of a Bacillus sphaericus binary toxin BinB. BMB Rep.
Bacillus sphaericus mosquito larvicidal toxin to cultured insect cells. Can. J.
43, 23–28.
Microbiol. 33, 982–989.
Broadwell, A.H., Baumann, P., 1986. Sporulation-associated activation of Bacillus
Davidson, E.W., Oei, C., Meyer, M., Bieber, A.L., Hindley, J., Berry, C., 1990. Interaction
sphaericus larvicide. Appl. Environ. Microbiol. 52, 758–764.
of the Bacillus sphaericus mosquito larvicidal proteins. Can. J. Microbiol. 36, 870–
Broadwell, A.H., Baumann, P., 1987. Proteolysis in the gut of mosquito larvae results
878.
in further activation of the Bacillus sphaericus toxin. Appl. Environ. Microbiol.
de Barjac, H., Veron, M., Cosmao Dumanoir, V., 1980. Caractérisation biochimique et
53, 1333–1337.
sérologique de souches de Bacillus sphaericus pathogènes ou non pour les
Broadwell, A.H., Baumann, L., Baumann, P., 1990a. Larvicidal properties of the 42
moustiques. Ann. Microbiol. (Inst. Pasteur). 131B, 191–201.
and 51 kilodalton Bacillus sphaericus proteins expressed in different bacterial
de Maagd, R.A., Bravo, A., Berry, C., Crickmore, N., Schnepf, H.E., 2003. Structure,
hosts: evidence for a binary toxin. Curr. Microbiol. 21, 361–366.
diversity, and evolution of protein toxins from spore-forming
Broadwell, A.H., Clark, M.A., Baumann, L., Baumann, P., 1990b. Construction by site-
entomopathogenic bacteria. Annu. Rev. Genet. 37, 409–433.
directed mutagenesis of a 39-kilodalton mosquitocidal protein similar to the
de Melo, J.V., Jones, G.W., Berry, C., Vasconcelos, R.H., Oliveira, C.M., Furtado, A.F.,
larva-processed toxin of Bacillus sphaericus 2362. J. Bacteriol. 172, 4032–4036.
Peixoto, C.A., Silva-Filha, M.H., 2009. Cry48Aa/Cry49Aa toxin from Bacillus
Brown, I.D., Watson, T.M., Carter, J., Purdie, D.M., Kay, B.H., 2004. Toxicity of
sphaericus displays cytopathological effects on susceptible and Binary toxin-
VectoLex (Bacillus sphaericus) products to selected Australian mosquito and
resistant Culex quinquefasciatus larvae. Appl. Environ. Microbiol. 75, 4782–4789.
nontarget species. J. Econ. Entomol. 97, 51–58.
de Oliveira, C.M., Filho, F.C., Beltran, J.E., Silva-Filha, M.H., Regis, L., 2003. Biological
Burke, W.F., McDonald, K.O., Davidson, E.W., 1983. Effect of UV light on spore
fitness of a Culex quinquefasciatus population and its resistance to Bacillus
viability and mosquito larvicidal activity of Bacillus sphaericus 1593. Appl.
sphaericus. J. Am. Mosquito Control Assoc. 19, 125–129.
Environ. Microbiol. 46, 954–956.
Des Rochers, B., Garcia, R., 1984. Evidence for persistence and recycling of Bacillus
Carboulec, N., Priest, F.G., 1989. Phenotypic characterisation of some strains of
sphaericus. Mosq. News 44, 160–165.
Bacillus sphaericus. Lett. Appl. Microbiol. 9, 113–116.
Elangovan, G., Shanmugavelu, M., Rajamohan, F., Dean, D.H., Jayaraman, K., 2000.
Carpusca, I., Jank, T., Aktories, K., 2006. Bacillus sphaericus mosquitocidal toxin
Identification of the functional site in the mosquito larvicidal binary toxin of
(MTX) and pierisin: the enigmatic offspring from the family of ADP-
Bacillus sphaericus 1593M by site-directed mutagenesis. Biochem. Biophys. Res.
ribosyltransferases. Mol. Microbiol. 62, 621–630.
Commun. 276, 1048–1055.
Chan, S.W., Thanabalu, T., Wee, B.Y., Porter, A.G., 1996. Unusual amino acid
El-Bendary, M., Priest, F.G., Charles, J.F., Mitchell, W.J., 2005. Crystal protein
determinants of host range in the Mtx2 family of mosquitocidal toxins. J. Biol.
synthesis is dependent on early sporulation gene expression in Bacillus
Chem. 271, 14183–14187.
sphaericus. FEMS Microbiol. Lett. 252, 51–56.
Charles, J.F., 1987. Ultrastructural midgut events in Culicidae larvae fed with Bacillus
From, C., Granum, P.E., Hardy, S.P., 2008. Demonstration of a cholesterol-dependent
sphaericus 2297 spore/crystal complex. Ann. Inst. Pasteur. 138, 471–484.
cytolysin in a noninsecticidal Bacillus sphaericus strain and evidence for
Charles, J.F., Nicolas, L., 1986. Recycling of Bacillus sphaericus 2362 in mosquito
widespread distribution of the toxin within the species. FEMS Microbiol. Lett.
larvae: a laboratory study. Ann. Inst. Pasteur. 137B, 101–111.
Gammon, K., Jones, G.W., Hope, S.J., Oliveira, C.M.F., Regis, L., Silva-Filha, M.H.,
Charles, J.F., Kalfon, A., Bourgouin, C., Debarjac, H., 1988. Bacillus sphaericus
Dancer, B.M., Berry, C., 2006. Conjugal transfer of a toxin-coding megaplasmid
asporogenous mutants – morphology, protein pattern and larvicidal activity.
from Bacillus thuringiensis subsp. israelensis to mosquitocidal strains of Bacillus
Ann. Inst. Pasteur. – Microbiol. 139, 243–259.
sphaericus. Appl. Environ. Microbiol. 73, 1766–1770.
Charles, J.-F., Silva-Filha, M.H., Nielsen-LeRoux, C., Humphreys, M.J., Berry, C., 1997.
Ge, Y., Hu, X., Zheng, D., Wu, Y., Yuan, Z., 2011. Allelic diversity and population
Binding of the 51- and 42-kDa individual components from the Bacillus
structure of Bacillus sphaericus revealed by multilocus sequence typing. Appl.
sphaericus crystal toxin to mosquito larval midgut membranes from Culex and
Environ. Microbiol.
Anopheles sp. (Diptera: Culicidae). FEMS Microbiol. Lett. 156, 153–159.
C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10 9

Hadapad, A.B., Vijayalakshmi, N., Hire, R.S., Dongre, T.K., 2008. Effect of ultraviolet Nicolas, L., Lecroisey, A., Charles, J.F., 1990. Role of the gut proteinases from
radiation on spore viability and mosquitocidal activity of an indigenous ISPC-8 mosquito larvae in the mechanism of action and the specificity of the Bacillus
Bacillus sphaericus Neide strain. Acta Trop. 107, 113–116. sphaericus toxin. Can. J. Microbiol. 36, 804–807.
Hadapad, A.B., Hire, R.S., Vijayalakshmi, N., Dongre, T.K., 2009. UV protectants for Nicolas, L., Nielsen-Leroux, C., Charles, J.F., Delecluse, A., 1993. Respective role of the
the biopesticide based on Bacillus sphaericus Neide and their role in protecting 42- and 51-kDa components of the Bacillus sphaericus toxin overexpressed in
the binary toxins from UV radiation. J. Invertebr. Pathol. 100, 147–152. Bacillus thuringiensis. FEMS Microbiol. Lett. 106, 275–280.
Han, B., Liu, H., Hu, X., Cai, Y., Zheng, D., Yuan, Z., 2007. Molecular characterization of Nicolas, L., Regis, L.N., Rios, E.M., 1994. Role of the exosporium in the stability of the
a glucokinase with broad hexose specificity from Bacillus sphaericus strain C3– Bacillus sphaericus binary toxin. FEMS Microbiol. Lett. 124, 271–275.
41. Appl. Environ. Microbiol. 73, 3581–3586. Nielsen-Leroux, C., Charles, J.-F., Thiéry, I., Georghiou, G.P., 1995. Resistance in a
Hayes, S.R., Hudon, M., Park, H.W., 2011. Isolation of novel Bacillus species showing laboratory population of Culex quinquefasciatus (Diptera: Culicidae) to Bacillus
high mosquitocidal activity against several mosquito species. J. Invertebr. sphaericus binary toxin is due to a change in the receptor on midgut brush-
Pathol. 107, 79–81. border membranes. Eur. J. Biochem. 228, 206–210.
Hazes, B., Read, R.J., 1995. A mosquitocidal toxin with a ricin-like cell-binding Nielsen-Leroux, C., Pasquier, F., Charles, J.-F., Sinègre, G., Gaven, B., Pasteur, N., 1997.
domain. Struct. Biol. 2, 358–359. Resistance to Bacillus sphaericus involves different mechanisms in Culex pipiens
Hindley, J., Berry, C., 1987. Identification, cloning and sequence analysis of the (Diptera: Culicidae) larvae. J. Med. Entomol. 34, 321–327.
Bacillus sphaericus 1593 41.9kD larvicidal toxin gene. Mol. Microbiol. 1, 187– Nielsen-LeRoux, C., Rao, D.R., Murphy, J.R., Carron, A., Mani, T.R., Hamon, S., Mulla,
194. M.S., 2001. Various levels of cross-resistance to Bacillus sphaericus strains in
Hire, R.S., Hadapad, A.B., Dongre, T.K., Kumar, V., 2009. Purification and Culex pipiens (Diptera: Culicidae) colonies resistant to B. sphaericus strain 2362.
characterization of mosquitocidal Bacillus sphaericus BinA protein. J. Invertebr. Appl. Environ. Microbiol. 67, 5049–5054.
Pathol. 101, 106–111. Nishiwaki, H., Nakashima, K., Ishida, C., Kawamura, T., Matsuda, K., 2007. Cloning,
Houman, F., Diaz-Torres, M.R., Wright, A., 1990. Transcriptional antitermination in functional characterization, and mode of action of a novel insecticidal pore-
the bgl operon of E. Coli is modulated by a specific RNA binding protein. Cell 62, forming toxin, Sphaericolysin, produced by Bacillus sphaericus. Appl. Environ.
1153–1163. Microbiol. 73, 3404–3411.
Hu, X., Fan, W., Han, B., Liu, H., Zheng, D., Li, Q., Dong, W., Yan, J., Gao, M., Berry, C., Oei, C., Hindley, J., Berry, C., 1990. An analysis of the genes encoding the 51.4- and
Yuan, Z., 2008. Complete genome sequences of the mosquitocidal bacterium 41.9-kDa toxins of Bacillus sphaericus 2297 by deletion mutagenesis: the
Bacillus sphaericus C3–41 and comparisons with closely related Bacillus species. J. construction of fusion proteins. FEMS Microbiol. Lett. 72, 265–274.
Bacteriol. 190, 2892–2902. Oei, C., Hindley, J., Berry, C., 1992. Binding of purified Bacillus sphaericus binary toxin
Humphreys, M.J., Berry, C., 1998. Variants of the Bacillus sphaericus binary toxins: and its deletion derivatives to Culex quinquefasciatus gut: elucidation of
implications for differential toxicity of strains. J. Invertebr. Pathol. 71, 184–185. functional binding domains. J. Gen. Microbiol. 138, 1515–1526.
Jones, G.W., Nielsen-Leroux, C., Yang, Y., Yuan, Z., Dumas, V.F., Monnerat, R.G., Berry, Oliveira, C.M., Silva-Filha, M.H., Nielsen-Leroux, C., Pei, G., Yuan, Z., Regis, L., 2004.
C., 2007. A new Cry toxin with a unique two-component dependency from Inheritance and mechanism of resistance to Bacillus sphaericus in Culex
Bacillus sphaericus. FASEB J. 21, 4112–4120. quinquefasciatus (Diptera: Culicidae) from China and Brazil. J. Med. Entomol.
Jones, G.W., Wirth, M.C., Monnerat, R.G., Berry, C., 2008. The Cry48Aa-Cry49Aa 41, 58–64.
binary toxin from Bacillus sphaericus exhibits highly-restricted target specificity. Oliveira, C.D., Tadei, W.P., Abdalla, F.C., 2009. Occurrence of apocrine secretion in
Environ. Microbiol. 10, 2418–2424. the larval gut epithelial cells of Aedes aegypti L., Anopheles albitarsis Lynch-
Karch, S., Monteny, N., Coz, J., 1988. Persistance of Bacillus sphaericus in a mosquito Arribalzaga and Culex quinquefasciatus say (Diptera: Culicidae): a defense
breeding site 4 years after its introduction for microbial control. CR Acad. Sci. strategy against infection by Bacillus sphaericus Neide? Neotrop. Entomol. 38,
Ser. II – Sci. Vie-Life Sci. 307, 289–292. 624–631.
Kellen, W.R., Clark, T.B., Lindegren, J.E., Ho, B.C., Rogoff, M.H., Singer, S., 1965. Opota, O., Charles, J.F., Warot, S., Pauron, D., Darboux, I., 2008. Identification and
Bacillus sphaericus Neide as a pathogen of mosquitoes. J. Invertebr. Pathol. 7, characterization of the receptor for the Bacillus sphaericus binary toxin in the
442–448. malaria vector mosquito, Anopheles gambiae. Comp. Biochem. Physiol. B –
Key, P.B., Scott, G.I., 1992. Acute toxicity of the mosquito larvicide, Bacillus Biochem. Mol. Biol. 149, 419–427.
sphaericus, to the grass shrimp, Palaemonetes pugio, and mummichog, Opota, O., Gauthier, N.C., Doye, A., Berry, C., Gounon, P., Lemichez, E., Pauron, D.,
Fundulus heteroclitus. Bull. Environ. Contam. Toxicol. 49, 425–430. 2011. Bacillus sphaericus binary toxin elicits host cell autophagy as a response to
Kramer, V.L., 1990. Efficacy and persistence of Bacillus sphaericus, Bacillus intoxication. PLoS One 6, e14682.
thuringiensis var. Israelensis, and methoprene against Culiseta incidens Park, H.-W., Bideshi, D., Federici, B.A., 2003. Recombinant strain of Bacillus
(Diptera: Culicidae) in tires. J. Econ. Entomol. 83, 1280–1285. thuringiensis producing Cyt1A, Cry11B, and the Bacillus sphaericus binary
Krych, V.K., Johnson, J.L., Yousten, A.A., 1980. Deoxyribonucleic acid homologies toxin. Appl. Environ. Microbiol. 69, 1331–1334.
among strains of Bacillus sphaericus. Int. J. Syst. Bacteriol. 30, 476–484. Partridge, M.R., Berry, C., 2002. Insecticidal activity of the Bacillus sphaericus Mtx1
Lacey, L.A., 1983. Larvicidal Activity of Bacillus Pathogens against Toxorhynchites toxin against Chironomus riparus. J. Invertebr. Pathol. 79, 135–136.
Mosquitos (Diptera, Culicidae). J. Med. Entomol. 20, 620–624. Pauchet, Y., Luton, F., Castella, C., Charles, J.F., Romey, G., Pauron, D., 2005. Effects of
Lacey, L.A., 2007. Bacillus thuringiensis serovariety israelensis and Bacillus sphaericus a mosquitocidal toxin on a mammalian epithelial cell line expressing its target
for mosquito control. J. Am. Mosquito Control Assoc. 23, 133–163. receptor. Cell. Microbiol. 7, 1335–1344.
Lacey, L.A., Lacey, C.M., Peacock, B., Thiery, I., 1988. Mosquito host range and field Pei, G., Oliveira, C.M.F., Yuan, Z., Nielsen-Leroux, C., Silva-Filha, M.H., Yan, J., Regis, L.,
activity of Bacillus sphaericus isolate 2297 (serotype 25). J. Am. Mosquito Control 2002. A strain of Bacillus sphaericus causes a slower development of resistance
Assoc. 4, 51–56. in Culex quinquefasciatus. Appl. Environ. Microbiol. 68, 3003–3009.
Liu, J.-W., Porter, A.G., Wee, B.Y., Thanabalu, T., 1996a. New gene from nine Bacillus Phannachet, K., Raksat, P., Limvuttegrijeerat, T., Promdonkoy, B., 2010. Production
sphaericus strains encoding highly conserved 35.8-kilodalton mosquitocidal and characterization of N- and C-terminally truncated Mtx2: a mosquitocidal
toxins. Appl. Environ. Microbiol. 62, 2174–2176. toxin from Bacillus sphaericus. Curr. Microbiol. 61, 549–553.
Liu, J.W., Yap, W.H., Thanabalu, T., Porter, A.G., 1996b. Efficient synthesis of Poncet, S., Delécluse, A., Anello, G., Klier, A., Rapoport, G., 1994. Transfer and
mosquitocidal toxins in Asticcacaulis excentricus demonstrates potential of expression of the cryIVB and cryIVD genes of Bacillus thuringiensis subsp.
gram-negative bacteria in mosquito control. Nat. Biotechnol. 14, 343–347. israelensis in Bacillus sphaericus 2297. FEMS Microbiol. Lett. 117, 91–95.
Marchler-Bauer, A., Lu, S., Anderson, J.B., Chitsaz, F., Derbyshire, M.K., DeWeese- Poncet, S., Bernard, C., Dervyn, E., Caylet, J., Klier, A., Rapoport, G., 1997.
Scott, C., Fong, J.H., Geer, L.Y., Geer, R.C., Gonzales, N.R., Gwadz, M., Hurwitz, D.I., Improvement of Bacillus sphaericus toxicity against dipteran larvae by
Jackson, J.D., Ke, Z., Lanczycki, C.J., Lu, F., Marchler, G.H., Mullokandov, M., integration, via homologous recombination, of the Cry11A toxin gene from
Omelchenko, M.V., Robertson, C.L., Song, J.S., Thanki, N., Yamashita, R.A., Zhang, Bacillus thuringiensis subsp. israelensis. Appl. Environ. Microbiol. 63, 4413–
D., Zhang, N., Zheng, C., Bryant, S.H., 2011. CDD: a conserved domain database 4420.
for the functional annotation of proteins. Nucl. Acids Res. 39, D225–D229. Priest, F.G., Ebdrup, L., Zahner, V., Carter, P.E., 1997. Distribution and
Matanmi, B.A., Federici, B.A., Mulla, M.S., 1990. Fate and persistence of Bacillus characterization of mosquitocidal toxin genes in some strains of Bacillus
sphaericus used as a mosquito larvicide in dairy wastewater lagoons. J. Am. sphaericus. Appl. Environ. Microbiol. 63, 1195–1198.
Mosquito Control Assoc. 6, 384–389. Promdonkoy, B., Promdonkoy, P., Wongtawan, B., Boonserm, P., Panyim, S., 2008.
Mathavan, S., Velpandi, A., Johnson, J.C., 1987. Sub-toxic effects of Bacillus sphaericus Cys31, Cys47, and Cys195 in BinA are essential for toxicity of a binary toxin
1593 M on feeding growth and reproduction of Laccotrephes griseus (Hemiptera: from Bacillus sphaericus. Curr. Microbiol. 56, 334–338.
Nepidae). Exp. Biol. 46, 149–153. Ramoska, W.A., Singer, S., Levy, R., 1977. Bioassay of three strains of Bacillus
Mulla, M.S., Thavara, U., Tawatsin, A., Chomposri, J., Su, T., 2003. Emergence of sphaericus on field-collected mosquito larvae. J. Invertebr. Pathol. 30, 151–
resistance and resistance management in field populations of tropical Culex 154.
quinquefasciatus to the microbial control agent Bacillus sphaericus. J. Am. Reinert, J.F., 2000. New classification for the composite genus Aedes (Diptera:
Mosquito Control Assoc. 19, 39–46. Culicidae: Aedini), elevation of subgenus Ochlerotatus to generic rank,
Mulligan, F.S., Schaefer, C.H., Miura, T., 1978. Laboratory and field evaluation of reclassification of the other subgenera, and notes on certain subgenera and
Bacillus sphaericus as a mosquito (Diptera-Culicidae) control agent. J. Econ. species. J. Am. Mosquito Control Assoc. 16, 175–188.
Entomol. 71, 774–777. Rippere, K.E., Johnson, J.L., Yousten, A.A., 1997. DNA similarities among mosquito-
Myers, P., Yousten, A., 1978. Toxic activity of Bacillus sphaericus SSII-1 for mosquito pathogenic and nonpathogenic strains of Bacillus sphaericus. Int. J. Syst.
larvae. Infect. Immun. 19, 1047–1053. Bacteriol. 47, 214–216.
Nicolas, L., Dossou-Yovo, J., Hougard, J.-M., 1987. Persistence and recycling of Rodcharoen, J., Mulla, M.S., 1997. Biological fitness of Culex quinquefasciatus
Bacillus sphaericus 2362 spores in Culex quinquefasciatus breeding sites in West (Diptera:Culicidae) susceptible and resistant to Bacillus sphaericus. J. Med.
Africa. Appl. Microbiol. Biotechnol. 25, 341–345. Entomol. 34, 5–10.
10 C. Berry / Journal of Invertebrate Pathology 109 (2012) 1–10

Romão, T.P., de Melo Chalegre, K.D., Key, S., Junqueira Ayres, C.F., Fontes de Oliveira, Israelensis by recombinant Caulobacter crescentus, a vehicle for biological control
C.M., de-Melo-Neto, O.P., Lobo Silva-Filha, M.H., 2006. A second independent of aquatic insect larvae. Appl. Environ. Microbiol. 58, 905–910.
resistance mechanism to Bacillus sphaericus binary toxin targets its alpha- Thanabalu, T., Berry, C., Hindley, J., 1993. Cytotoxicity and ADP-ribosylating activity
glucosidase receptor in Culex quinquefasciatus. FEBS J. 273, 1556–1568. of the mosquitocidal toxin from Bacillus sphaericus SSII-1: possible roles of the
Romao, T.P., de-Melo-Neto, O.P., Silva-Filha, M.H., 2011. The N-terminal third of the 27- and 70-kilodalton peptides. J. Bacteriol. 175, 2314–2320.
BinB subunit from Bacillus sphaericus binary toxin is sufficient for its interaction Tietze, N.S., Olson, M.A., Hester, P.G., Moore, J.J., 1993. Tolerance of sewage
with midgut receptors in Culex quinquefasciatus. FEMS Microbiol. Lett. treatment plant microorganisms to mosquitocides. J. Am. Mosquito Control
Rungrod, A., Tjahaja, N.K., Soonsanga, S., Audtho, M., Promdonkoy, B., 2009. Bacillus Assoc. 9, 477–479.
sphaericus Mtx1 and Mtx2 toxins co-expressed in Escherichia coli are synergistic Treiber, N., Reinert, D.J., Carpusca, I., Aktories, K., Schulz, G.E., 2008. Structure and
against Aedes aegypti larvae. Biotechnol. Lett. 31, 551–555. mode of action of a mosquitocidal holotoxin. J. Mol. Biol. 381, 150–159.
Sanitt, P., Promdonkoy, B., Boonserm, P., 2008. Targeted mutagenesis at charged Watanabe, M., Enomoto, S., Takamura-Enya, T., Nakano, T., Koyama, K., Sugimura, T.,
residues in Bacillus sphaericus BinA toxin affects mosquito-larvicidal activity. Wakabayashi, K., 2004. Enzymatic properties of pierisin-1 and its N-terminal
Curr. Microbiol. 57, 230–234. domain, a guanine-specific ADP-ribosyltransferase from the cabbage butterfly.
Savage, H.M., Strickman, D., 2004. The genus and subgenus categories within J. Biochem. 135, 471–477.
Culicidae and placement of Ochlerotatus as a subgenus of Aedes. J. Am. Mosquito Wati, M., Thanabalu, T., Porter, A., 1997. Gene from tropical Bacillus sphaericus
Control Assoc. 20, 208–214. encoding a protease closely related to subtilisins from Antarctic bacilli. Biochim.
Schirmer, J., Just, I., Aktories, K., 2002a. The ADP-ribosylating mosquitocidal toxin Biophys. Acta 1352, 56–62.
from Bacillus sphaericus. J. Biol. Chem. 277, 11941–11948. White, P.J., Lotay, H.K., 1980. Minimal nutritional requirements of Bacillus sphaericus
Schirmer, J., Wieden, H.-J., Rodnina, M.V., Aktories, K., 2002b. Inactivation of the NCTC9602 and 26 other strains of this species: the majority grow and sporulate
elongation factor Tu by mosquitocidal toxin-catalysed mono-ADP-ribosylation. with acetate as sole major source of carbon. J. Gen. Microbiol. 118, 13–19.
Appl. Environ. Microbiol. 68, 4894–4899. Wirth, M.C., Federici, B.A., Walton, W.E., 2000. Cyt1A from Bacillus thuringiensis
Schwartz, J.-L., Potvin, L., Coux, F., Charles, J.-F., Berry, C., Humphreys, M.J., Jones, synergizes activity of Bacillus sphaericus against Aedes aegypti (Diptera:
A.F., Bernhart, I., Dalla Serra, M., Menestrina, G., 2001. Permeabilization of Culicidae). Appl. Environ. Microbiol. 66, 1093–1097.
model lipid membranes by Bacillus sphaericus mosquitocidal binary toxin and Wirth, M.C., Georghiou, G.P., Malik, J.I., Abro, G.H., 2000b. Laboratory selection for
its individual components. J. Membrane Biol. 184, 171–183. resistance to Bacillus sphaericus in Culex quinquefasciatus (Diptera: Culicidae)
Shanmugavelu, M., Rajamohan, F., Kathirvel, M., Elangovan, G., Dean, D.H., from California, USA. J. Med. Entomol. 37, 534–540.
Jayaraman, K., 1998. Functional complementation of nontoxic mutant binary Wirth, M.C., Delécluse, A., Walton, W.E., 2001. Cyt1Ab1 and Cyt2Ba1 from Bacillus
toxins of Bacillus sphaericus 1593M generated by site-directed mutagenesis. thuringiensis subsp. Medellin and B. thuringiensis subsp. israelensis synergize
Appl. Environ. Microbiol. 64, 756–759. Bacillus sphaericus against Aedes aegypti and resistant Culex quinquefasciatus
Silapanuntakul, S., Pantuwatana, S., Bhumiratana, A., Chaaroensiri, K., 1983. The (Dipetera: Culicidae). Appl. Environ. Microbiol. 67, 3280–3284.
comparative persistence of toxicity of Bacillus sphaericus strain 1593 and Wirth, M.C., Jiannino, J.A., Federici, B.A., Walton, W.E., 2004. Synergy between toxins
Bacillus thuringiensis serotype H-14 against mosquito larvae in different kinds of of Bacillus thuringiensis subsp. Israelensis and Bacillus sphaericus. J. Med.
environments. J. Invertebr. Pathol. 42, 387–392. Entomol. 41, 935–941.
Silva-Filha, M.H., Nielsen-LeRoux, C., Charles, J.-F., 1997. Binding kinetics of Bacillus Wirth, M.C., Yang, Y., Walton, W.E., Federici, B.A., Berry, C., 2007. Mtx toxins
sphaericus binary toxin to midgut brush-border membranes of Anopheles and synergize Bacillus sphaericus and Cry11Aa against susceptible and insecticide-
Culex sp. Mosquito larvae. Eur. J. Biochem. 247, 754–761. resistant Culex quinquefasciatus. Appl. Environ. Microbiol. 73, 6066–6071.
Silva-Filha, M.H., Nielsen-LeRoux, C., Charles, J.-F., 1999. Identification of the Wraight, S.P., Molloy, D.P., Singer, S., 1987. Studies on the culicine mosquito host
receptor for Bacillus sphaericus crystal toxin in the brush border membrane of range of Bacillus sphaericus and Bacillus thuringiensis var. Israelensis with notes
the mosquito Culex pipiens (Diptera: Culicidae). Insect Biochem. Mol. Biol. 29, on the effects of temperature and instar on bacterial efficacy. J. Invertebr.
711–721. Pathol. 49, 291–302.
Singer, S., 1973. Insecticidal activity of recent bacterial isolates and their toxins Wu, E., Jun, L., Yuan, Y., Yan, J., Berry, C., Yuan, Z., 2007. Characterization of a cryptic
against mosquito larvae. Nature 244, 110–111. plasmid from Bacillus sphaericus strain LP1-G. Plasmid 57, 296–305.
Singh, G.J., Gill, S.S., 1988. An electron microscope study of the toxic action of Xudong, X., Renqiu, K., Yuxiang, H., 1993. High larvicidal activity of intact
Bacillus sphaericus in Culex quinquefasciatus larvae. J. Invertebr. Pathol. 52, 237– recombinant cyanobacterium Anabaena sp. PCC7120 expressing gene 51 and
247. gene 42 of Bacillus sphaericus sp. 2297. FEMS Microbiol. Lett. 107, 247–250.
Singkhamanan, K., Promdonkoy, B., Chaisri, U., Boonserm, P., 2010. Identification of Yang, Y., Wang, L., Gaviria, A., Yuan, Z., Berry, C., 2007. Proteolytic stability of
amino acids required for receptor binding and toxicity of the Bacillus sphaericus insecticidal toxins expressed in recombinant bacilli. Appl. Environ. Microbiol.
binary toxin. FEMS Microbiol. Lett. 303, 84–91. 73, 218–225.
Smith, A.W., Camara-Artigas, A., Allen, J.P., 2004. Crystallization of the mosquito- Yap, W.H., Thanabalu, T., Porter, A.G., 1994. Expression of mosquitocidal toxin genes
larvicidal binary toxin produced by Bacillus sphaericus. Acta Crystallogr. 60, in a gas-vacuolated strain of Ancylobacter aquaticus. Appl. Environ. Microbiol.
952–953. 60, 4199–4202.
Smith, A.W., Camara-Artigas, A., Brune, D.C., Allen, J.P., 2005. Implications of high- Young, M.D., Undeen, A.H., Dame, D.A., Wing, S.R., 1990. The effect of Bacillus
molecular-weight oligomers of the binary toxin from Bacillus sphaericus. J. sphaericus upon the susceptibility of Anopheles quadrimaculatus to Plasmodium
Invertebr. Pathol. 88, 27–33. berghei. J. Am. Mosquito Control Assoc. 6, 139–140.
Stray, J.E., Klowden, M.J., Hurlbert, R.E., 1988. Toxicity of Bacillus sphaericus crystal Yousten, A.A., 1984. Bacteriophage typing of mosquito pathogenic strains of Bacillus
toxin to adult mosquitoes. Appl. Environ. Microbiol. 54, 2320–2321. sphaericus. J. Invertebr. Pathol. 43, 124–125.
Tanapongpipat, S., Nantapong, N., Cole, J., Panyim, S., 2003. Stable integration and Yousten, A.A., Fretz, S.B., Jelley, S.A., 1985. Selective medium for mosquito-
expression of mosquito-larvicidal genes from Bacillus thuringiensis subsp. pathogenic strains of Bacillus sphaericus. Appl. Environ. Microbiol. 49, 1532–
Israelensis and Bacillus sphaericus into the chromosome of Enterobacter 1533.
amnigenus: a potential breakthrough in mosquito biocontrol. FEMS Microbiol. Yousten, A.A., Benfield, E.F., Campbell, R.P., Foss, S.S., Genthner, F.J., 1991. Fate of
Lett. 221, 243–248. Bacillus sphaericus 2362 spores following ingestion by nontarget invertebrates.
Tandeau de Marsac, N., de la Torre, F., Szulmajster, J., 1987. Expression of the J. Invertebr. Pathol. 58, 427–435.
larvicidal gene of Bacillus sphaericus 1593M in the cyanobacterium Anacystis Yousten, A.A., Genthner, F.J., Benfield, E.F., 1992. Fate of Bacillus sphaericus and
nidulans R2. Mol. Gen. Genet. 209, 396–398. Bacillus thuringiensis serovar israelensis in the aquatic environment. J. Am.
Tangsongcharoen, C., Boonserm, P., Promdonkoy, B., 2011. Functional Mosquito Control Assoc. 8, 143–148.
characterization of truncated fragments of Bacillus sphaericus binary toxin Yuan, Z., Rang, C., Maroun, R.C., Juarez-Perez, V., Frutos, R., Pasteur, N., Vendrely, C.,
BinB. J. Invertebr. Pathol. 106, 230–235. Charles, J.F., Nielsen-Leroux, C., 2001. Identification and molecular structural
Taylor, L.D., Burke, W.F., 1990. Transformation of an entomopathic strain of prediction analysis of a toxicity determinant in the Bacillus sphaericus crystal
Bacillus sphaericus by high-voltage electroporation. FEMS Microbiol. Lett. 66, larvicidal toxin. Eur. J. Biochem. 268, 2751–2760.
125–127. Yuan, Z.M., Pei, G.F., Regis, L., Nielsen-Leroux, C., Cai, Q.X., 2003. Cross-resistance
Thanabalu, T., Porter, A.G., 1995. Efficient expression of a 100-kilodalton between strains of Bacillus sphaericus but not B. Thuringiensis israelensis in
mosquitocidal toxin in protease-deficient recombinant Bacillus sphaericus. colonies of the mosquito Culex quinquefasciatus. Med. Vet. Entomol. 17, 251–
Appl. Environ. Microbiol. 61, 4031–4036. 256.
Thanabalu, T., Porter, A.G., 1996. A Bacillus sphaericus gene encoding a novel type of Zahiri, N.S., Mulla, M.S., 2005. Non-larvicidal effects of Bacillus thuringiensis
mosquitocidal toxin of 31.8 kDa. Gene 170, 85–89. israelensis and Bacillus sphaericus on oviposition and adult mortality of Culex
Thanabalu, T., Hindley, J., Jackson-Yap, J., Berry, C., 1991. Cloning, sequencing and quinquefasciatus Say (Diptera: Culicidae). J. Vector Ecol. 30, 155–162.
expression of a gene encoding a 100-kilodalton mosquitocidal toxin from Zahner, V., Priest, F.G., 1997. Distribution of restriction endonucleases among some
Bacillus sphaericus SSII-1. J. Bacteriol. 173, 2776–2785. entomopathogenic strains of Bacillus sphaericus. Lett. Appl. Microbiol. 24, 483–
Thanabalu, T., Hindley, J., Berry, C., 1992a. Proteolytic processing of the 487.
mosquitocidal toxin from Bacillus sphaericus SSII-1. J. Bacteriol. 174, 5051–5056. Zahner, V., Momen, H., Priest, F.G., 1998. Serotype H5a5b is a major clone within
Thanabalu, T., Hindley, J., Brenner, S., Oei, C., Berry, C., 1992b. Expression of the mosquito-pathogenic strains of Bacillus sphaericus. Syst. Appl. Microl. 21, 162–
mosquitocidal toxins of Bacillus sphaericus and Bacillus thuringiensis subsp. 170.

You might also like