You are on page 1of 24

Molecular Ecology (2014) 23, 1473–1496 doi: 10.1111/mec.

12421

SPECIAL ISSUE: NATURE’S MICROBIOME


INVITED REVIEWS AND META-ANALYSES
The impact of microbial symbionts on host plant
utilization by herbivorous insects
A L L I S O N K . H A N S E N and N A N C Y A . M O R A N
Yale University, West Haven, CT 06516-7388, USA

Abstract
Herbivory, defined as feeding on live plant tissues, is characteristic of highly successful
and diverse groups of insects and represents an evolutionarily derived mode of feeding.
Plants present various nutritional and defensive barriers against herbivory; neverthe-
less, insects have evolved a diverse array of mechanisms that enable them to feed and
develop on live plant tissues. For decades, it has been suggested that insect-associated
microbes may facilitate host plant use, and new molecular methodologies offer the pos-
sibility to elucidate such roles. Based on genomic data, specialized feeding on phloem
and xylem sap is highly dependent on nutrient provisioning by intracellular symbionts,
as exemplified by Buchnera in aphids, although it is unclear whether such symbionts
play a substantive role in host plant specificity of their hosts. Microorganisms present
in the gut or outside the insect body could provide more functions including digestion
of plant polymers and detoxification of plant-produced toxins. However, the extent of
such contributions to insect herbivory remains unclear. We propose that the potential
functions of microbial symbionts in facilitating or restricting the use of host plants are
constrained by their location (intracellular, gut or environmental), and by the fidelity of
their associations with insect host lineages. Studies in the next decade, using molecular
methods from environmental microbiology and genomics, will provide a more compre-
hensive picture of the role of microbial symbionts in insect herbivory.
Keywords: Buchnera, coevolution, gut microbiota, insect nutrition, insect–plant–microbe inter-
action, intracellular symbiont

Received 15 April 2013; revision received 2 June 2013; accepted 12 June 2013

herbivory is a successful feeding mode, but only after


Introduction
key hurdles are overcome.
Herbivory, defined here as feeding on live plant tissues, Obvious obstacles to herbivory include the low nutri-
is an evolutionarily derived mode of feeding in insects ent content, indigestibility and toxicity of many plant
(Strong & Lawton 1984; Labandeira & Sepkoski 1993). tissues. Although plants represent the largest source of
Evolutionary transitions to herbivory, specifically biomass in terrestrial systems, their tissues are largely
herbivory on angiosperms, have been followed by accel- composed of recalcitrant molecules such as cellulose,
erated rates of insect species diversification (Mitter et al. hemicellulose, pectin and lignin that are difficult or
1988; Farrell & Farrell 1998). Herbivorous insects are impossible to digest for most animals (Watanabe &
indeed extraordinarily diverse, as measured both by Tokuda 2010). Furthermore, the levels of nitrogen rela-
species numbers and by their varied lifestyles (Strong & tive to carbon tend to be low when compared to the
Lawton 1984; May 1988; Farrell 1998; Futuyma & Agra- nutritional needs of animals and exhibit substantial var-
wal 2009). This macroevolutionary pattern implies that iation among different plant tissues, growth stages and
species (Mattson 1980). Besides being generally nitrogen
Correspondence: Allison K. Hansen, Fax: 203-737-3109; poor, some plant tissues present very unbalanced pro-
E-mail: akh@illinois.edu; and Nancy A. Moran, files of amino acids, so that growth may be severely
Fax: 203-737-3109; E-mail: nancy.moran@austin.utexas.edu limited by shortages of particular essential amino acids,

© 2013 John Wiley & Sons Ltd


1474 A . K . H A N S E N and N . A . M O R A N

or of vitamins (e.g. Sandstr€ om & Moran 1999). Nutritive bacteria and fungi, possess many functional capabilities
quality is often lowered further because many plants entirely absent from animals. These include biosynthetic
contain compounds such as protease inhibitors that capabilities for universally required metabolites such as
inactivate digestive enzymes in the insect gut, prevent- vitamins and some amino acids, catabolic capabilities
ing digestion and absorption (Ryan 1990). Finally, plant such as cellulolytic activity and production of many
tissues are laced with a wide variety of other toxic allel- toxic molecules. As such, they present a vast new tool
ochemicals that vary extensively among plant taxa and kit with potential to facilitate herbivory as a lifestyle.
in mode of action on their target. There is a vast litera- On the other hand, symbiont acquisition by hosts may
ture on the many morphological, metabolic and be unreliable, or tissue localization of symbionts may
behavioural adaptations of insects that enable herbivory limit their contributions to specific host functions.
and on the adaptations of plants for avoiding negative Studies of herbivore–plant interactions span different
impacts of insect herbivory (Schoonhoven et al. 2005). levels of analysis, from mechanistic studies of how par-
Recently, more researchers have begun to appreciate ticular insects digest specific plant tissues to evolution-
the potential for associations with symbiotic microbes as ary and population analyses. We consider all of these,
a means of expanding ecological range of eukaryotic with emphasis on the use of new genomic data and
hosts and of animal hosts in particular (Moran 2007; molecular techniques to infer symbiont roles in deter-
McFall-Ngai et al. 2013). Symbionts, defined as microbes mining abilities to overcome obstacles to feeding on
that form persistent, noninvasive associations with hosts, particular plant tissues or species. This evaluation nec-
are common in insects generally and in herbivorous essarily involves considering specific mechanisms
insects in particular. Microbes offer capabilities for syn- through which symbiotic microorganisms might enable
thesizing nutrients, digesting recalcitrant plant polymers use of particular plant tissues or plant species.
and neutralizing plant toxins. The possibility that symbi-
otic associations could shape interactions with host plants
Broad categories of symbionts
has been proposed by numerous researchers (reviewed
in Berenbaum 1988; Barbosa et al. 1991; Bernays & Chap- For purposes of this review, we define symbiont
man 1994; Felton & Tumlinson 2008; Janson et al. 2008; broadly to include any microbes that have persistent
Clark et al. 2010; Feldhaar 2011; Frago et al. 2012). How- associations with their hosts and that do not cause overt
ever, the extent to which insect symbionts impact their pathogenic symptoms. Insect symbionts can be divided
insect host’s ability to feed on plant tissues or shape spe- into three general categories based on their location
cific interactions between herbivorous insects and partic- relative to the insect body (Fig. 1). First, some symbio-
ular plant species remains an open question. nts are intracellular or housed within the body cavity
In this review, we evaluate the impact of symbionts and are passed between generations through eggs,
on whether insects can use plants as food and on which resulting in maternal inheritance. Heritable, usually
plants are used, and how. The primary reason for intracellular, symbionts are found in a variety of insect
expecting symbionts to be able to assist insects in groups, which often evolve specialized cells (bacterio-
exploiting plant hosts is that microorganisms, including cytes) for housing them. Second, microbes inhabit the

Fig. 1 Potential contributions of microor-


ganisms to host ability to feed on live
plants are dependent on the symbiont
location inside or outside the insect
body.

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1475

guts of most animals including most insects, living from competition from other microbes and require only
within the gut lumen where they form communities of simple host-derived carbon and nitrogen inputs for
varying complexity in different insect species. Third, energy. In contrast, symbionts that digest complex plant
microbes regularly colonize food before it is ingested metabolites into absorbable nitrogen or energy sources
and may impact its composition. These environmental are extracellular, either in the lumen of the gut or out-
symbionts most likely are important in insects that feed side the host’s body, where complex macromolecule
on a stored food resource within a nest or localized breakdown must occur before absorption. In this
feeding area such as a gall or gallery. section, we will discuss current knowledge of how
The potential functions of microbial symbionts in intracellular insect symbionts provision their herbivore
facilitating or restricting the use of host plants are con- hosts with required nutrients lacking in the diet and the
strained by their location and by the persistence of implications for how these symbionts allow insect
their associations with host lineages. For example, heri- herbivores to exploit living plants as food.
table symbionts form long-term associations with host Intracellular nutritional symbionts that are inherited
lineages and are thus ideally positioned to contribute maternally are found in a variety of insect taxa, includ-
evolutionarily significant capabilities, but they are often ing many herbivorous groups. For example, two of the
confined to the cytoplasm of specialized host cells and most diverse insect groups, the leaf beetles (Chrysome-
thus unable to secrete gene products directly into the lidae) and the weevils (Curculionoidea), include species
gut lumen or into oral secretions, where they might containing bacterial symbionts that are transmitted
aid in digestion, detoxification or modification of host through eggs and believed to be critical for nutrition of
plant recognition. Gut symbionts or environmental hosts (Conord et al. 2008; K€ olsch and Pedersen 2010).
symbionts are better positioned for contributing these Symbioses of sap-feeding insects include the most
functions, but their transmission among individual intensively studied cases, so we focus on those in this
hosts may be less dependable in some insect life section.
cycles, so their contributions might often be too unreli- Obligate intracellular symbioses are universal in
able to facilitate the expansion of host plant range. hemipteran insects that feed on phloem or xylem sap;
Intracellular symbionts can contribute small molecules these include most members of the suborders Stern-
to host metabolism, but these contributions will orrhyncha (aphids, whiteflies, psyllids, scale insects)
depend on host capabilities for transporting these and of the Auchenorrhyncha (leafhoppers, cicadas, spit-
molecules across membranes and throughout the insect tlebugs and planthoppers) (Buchner 1965; Douglas 1998;
body. Moran et al. 2008; Urban & Cryan 2012). Obligate
Finally, these three symbiont categories differ in symbionts are also widespread in hemipteran bugs
whether they have dynamic genomes that can acquire feeding on plant cell contents, such as stink bugs and
new abilities as they evolve with hosts. In particular, plant bugs (suborder Heteroptera, including Coreidae,
genomes of heritable symbionts typically undergo Alydidae, Pentatomidae, Plataspidae) (e.g. (Hosokawa
extensive gene loss over their long-term evolution and et al. 2006)). In many of these systems, phylogenetic
do not incorporate novel genes, leading to dwindling studies have revealed an evolutionary history of long-
metabolic capabilities (McCutcheon & Moran 2012). If term codiversification of insect hosts with bacterial sym-
herbivorous insects need to respond to both short-term bionts, with the associations dating to the origins of
and long-term changes in the nutritional quality, plant major groups such as aphids and psyllids and in some
recognition or toxicity of plant tissues that they ingest, cases dating to the earliest stages of terrestrial herbiv-
long-term obligate symbionts with reduced genomes ory, on Permian vascular plants at least 270 million
will be unlikely to provide the necessary novelty. In years ago (Munson et al. 1991; Thao et al. 2000a,b; Bau-
contrast, most facultative heritable symbionts, as well as mann 2005; Downie & Gullan 2005; Moran et al. 2005;
gut and environmental symbionts show dynamic Moran 2007). In most cases, obligate intracellular symbi-
genomes, featuring ongoing gain and loss of functional onts are bacteria, and the symbiotic lifestyle has arisen
genes, genome rearrangements and mobile elements, multiple times within several different bacterial groups,
processes that are typical of most bacterial lineages. including Gammaproteobacteria, Betaproteobacteria, Alpha-
proteobacteria and Bacteroidetes (Fig. 2, Moran et al. 2008).
Intracellular heritable symbionts
Widespread herbivore–symbiont expansions into
Many insect symbionts that synthesize nutrients for
sap-feeding niches
their insect hosts are intracellular and obligate for the
insect’s survival and reproduction. Intracellular symbio- The central basis to these obligate symbioses is nutri-
nts live in a stable environment generally protected ent provisioning, as revealed by both genomic and

© 2013 John Wiley & Sons Ltd


1476 A . K . H A N S E N and N . A . M O R A N

Genome Val
Insect species Specialization Nutritional symbiont size (kb) Orn Arg His Lys Thr BCA Ile Leu Chorismate Trp Phe Met Sul Cys
Heteropsylla texana Ph γ Carsonella ruddii HT 158
γ Carsonella ruddii HC & 166
Heteropsylla cubana Ph γ SS HC 1,122

Ctenarytaina spatulata Ph γ Carsonella ruddii CS 163

γ Carsonella ruddii CE & 163


Ctenarytaina eucalypti Ph γ SS CE 1,441

Pachypsylla venusta Ph γ Carsonella ruddii PV 160

Pachypsylla celtidis Ph γ Carsonella ruddii PC 160

Bemisia tabaci Ph γ Portiera aleyrodidarum 358

Acyrthosiphon pisum Ph γ Buchnera aphidicola APS 656

Acyrthosiphon kondoi Ph γ Buchnera aphidicola AK 653

Uroleucon ambrosiae Ph γ Buchnera aphidicola UA 628

Schizaphis graminum Ph γ Buchnera aphidicola Sg 653

Ph
Hemiptera

Baizongia pistaciae γ Buchnera aphidicola Bp 618

γ Buchnera aphidicola Cc & 425


Cinara cedri Ph γ Serratia symbiotica Cc 1,763

Cinara tujafilina Ph γ Buchnera aphidicola Ct 453

Aspidiotus nerii Me F Uzinura diaspidicola 263

β Tremblaya princeps & 139


Planococcus citri Ph γ Moranella endobia 538

Homalodisca vitripennis Xy F Sulcia muelleri Hc & 246


γ Baumannia cicadellinicola 686

F Sulcia muelleri DSEM & 277


Xy
*
Diceroprocta semicincta α Hodgkinia cicadicola 144
F Sulcia muelleri CARI & 277
Clastoptera arizonana Xy
β Zinderia insecticola 209

Megacopta punctatissima Ph γ Ishikawaella capsulata 746

= long-term obligate symbiont encodes gene = Gene not encoded


= obligate co-symbiont encodes gene = Gene not required because
alternative pathway encoded
= putative insect homolog complements
pathway

Fig. 2 Amino acid biosynthesis capabilities of long-term nutritional symbionts with fully sequenced genomes for herbivorous insects
in the Hemiptera. Evolutionary relationships are based on Cryan & Urban (2012). Blue branches represent Sternorrhyncha, orange
branches represent Auchenorrhyncha and the purple branch represents Heteroptera. Specialization refers to the plant tissue eaten
(Ph, phloem; Me, mesophyll and Xy, xylem). Bacterial symbiont taxa symbols represent c, Gammaproteobacteria; b, Betaproteobacte-
ria; a, Alphaproteobacteria; F, Flavobacteriales. Genes from essential amino acid pathways and pathways of their intermediates are
presented as rectangles to the right of the insect tree. Orn, ornithine; Arg, Arginine; His, histidine; Lys, lysine; Thr, threonine; BCA,
shared branched-chain amino acid pathway; Val, valine; Ile, isoleucine; Leu, leucine; chorismate, shared pathway for aromatic amino
acids; Trp, tryptophan; Phe, phenylalanine; Met, methionine pathway (starting from the homoserine intermediate); Sul, sulphur
reduction and Cys, cysteine. *alternative methionine synthetase (MetH) requires additional genes for biosynthesis of a B12-like cofac-
tor; these are encoded in the Hodgkinia genome but not shown. Information for gene presence is based on GenBank data, and rele-
vant papers are cited in the main text.

experimental studies (Douglas 1998; Moran et al. 2008; with intracellular, heritable bacteria that provision
Hansen & Moran 2011). Animals feeding on plants are essential amino acids (Mittler 1971; Douglas 1988;
generally limited by available nitrogen (Schoonhoven Douglas & Prosser 1992; Febvay et al. 1995; Shigenobu
et al. 2005), and nitrogen concentrations are typically et al. 2000; Baumann 2005). Given the nutritional
two to ten times lower in phloem and xylem, respec- requirements of insects, the metabolic capabilities of
tively, compared with leaves (Mattson 1980). Further- bacteria and the restricted nutrient content of xylem
more, xylem and phloem sap contain unbalanced and phloem sap, it is no surprise that nutrition-provi-
profiles of amino acids and are especially deficient in sioning symbionts are prevalent among sap-feeding
the so-called essential amino acids, that is, those that insect taxa (Fig. 2).
cannot be synthesized by animals (Slansky & Scriber In these ancient obligate symbioses, a consequence of
1985; Andersen et al. 1989; Sandstr€ om & Pettersson the long-term maternal transmission and resulting
1994; Sandstr€ om & Moran 2001). Sap-feeding insects clonality of symbionts is that their genomes are drasti-
overcome these nutritional limitations by associating cally reduced in size and gene number and display

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1477

rapid sequence evolution and an A+T bias in nucleotide mechanisms for generating their own carotenoids, some-
base composition (McCutcheon & Moran 2012). Despite times through incorporation of foreign genes directly into
loss of up to 90% of ancestral genes, all of these obligate the insect genome (see Box 2). In the case of whiteflies,
symbiont genomes retain genes for essential amino acid carotenoid biosynthetic genes are intact in the genome of
pathways but have lost most genes for nonessential their obligate symbiont (Portiera), despite its extremely
amino acid biosynthesis, consistent with their nutri- reduced genome (358 kb) (Sloan & Moran 2012b).
tional roles (Shigenobu et al. 2000; Moran et al. 2008; In sum, heritable, intracellular symbionts have been
McCutcheon & Moran 2012; Fig. 2). While a few genes critical in enabling insects to occupy and dominate
encoding enzymes in certain essential amino acid path- ecological niches in which essential amino acids are
ways are sometimes missing (Fig. 2), genes present in limiting, such as dietary specialization on xylem and
the insect host and/or other co-occurring obligate sym- phloem sap (Buchner 1965; Baumann 2005; Moran et al.
bionts appear to provide these missing enzymes 2005; Moran 2007).
(Fig. 2). For example, co-occurring obligate symbionts
in six independently evolved symbiont systems encode
Obligate symbiont genomes and their influence on host
enzymes that complement enzymes or pathways miss-
plant breadth
ing from their partner symbiont (Perez-Brocal et al.
2006; Wu et al. 2006; Gosalbes et al. 2008; McCutcheon Insect adaptations involving morphology, behaviour
& Moran 2007; McCutcheon et al. 2009; McCutcheon & and physiology have enabled and restricted insect use
Moran 2010; McCutcheon & von Dohlen 2011; Sloan & of particular host plants (Bernays & Chapman 1994;
Moran 2012a; Fig. 2). Schoonhoven et al. 2005). As discussed above, heritable
In addition to these genomic data, experimental symbionts enable insects to utilize phloem and xylem
results available for the aphid-Buchnera system also sup- sap as food; however, it is not clear whether these sym-
port a role of obligate symbionts in amino acid provi- bionts are major determinants of host plant range. Sym-
sioning (Douglas 1998; Akman G€ und€ uz & Douglas biont-driven host plant specificity would depend on
2009; Douglas et al. 2011; Macdonald et al. 2011). Buch- differences among plants in nutrient profiles, because
nera uses nonessential amino acids as substrates for the the main role of obligate symbionts is to synthesize and
production of essential amino acids. Specifically, Buch- provision essential nutrients, particularly amino acids
nera may efficiently use glutamate synthesized by insect and vitamins. Potentially, obligate symbionts can con-
enzymatic machinery (GOGAT cycle), which recycles tribute to host plant specificity by provisioning nutri-
waste ammonia, to sustain this nitrogen-limited mutual- ents lacking in specific plant species, thereby allowing
ism (Hansen & Moran 2011), although the extent of an insect to develop and reproduce on a plant other-
such recycling is not yet clear (Macdonald et al. 2012). wise nutritionally unsuitable. Alternatively, in addition
Although amino acids generally appear to be the to the nutritional capacity of the obligate symbiont,
primary nutrients provided by obligate symbionts in acquisition of an additional cosymbiont could buffer the
sap-feeding insects, genome sequencing suggests that insect from host plant-derived nutrient deficiencies,
these symbionts likely also provision cofactors and expanding its host plant range. Conversely, symbionts
other potentially beneficial compounds to hosts. In pea that lose particular biosynthetic pathways may con-
aphids (Acyrthosiphon pisum), Buchnera encodes genes strain insect hosts to plant species or tissue types that
involved in riboflavin production (Shigenobu et al. are enriched in these missing nutrients (Fig. 3).
2000). Baumannia, one of the cosymbionts of the sharp- Comparative analyses of genomes from different obli-
shooter Homalodisca vitripennis (Cicadellidae: Cicadelli- gate symbiont lineages of diverse hemipteran taxa
nae), retains 84 genes predicted to enable the reveal convergent patterns in gene loss and retention
production of several B-vitamins (Wu et al. 2006). (Moran et al. 2008; McCutcheon & Moran 2012). When
Baumannia was the first obligate symbiont sequenced examining symbionts across sap-feeding hemipteran
from a xylem-feeding host, and the presence of these species, with single or partner symbionts, the same
genes was initially hypothesized to reflect the absence overall pattern is evident: symbionts are able to provi-
of these compounds from the xylem sap diet of sharp- sion most or all essential amino acids in each insect
shooters. However, subsequent genome sequences of host species (Fig. 2). This similarity reflects the known
symbionts of xylem feeders lack many of these genes deficiency of essential amino acids in both phloem sap
(McCutcheon & Moran 2012). Carotenoids also appear and xylem sap across diverse plants (Andersen et al.
to be beneficial or required nutrients for phloem-feed- 1989; Andersen et al. 1992; Brodbeck et al. 1990; Sands-
ing insects, and phloem lacks these compounds, which tr€
om & Pettersson 1994). Because the single or paired
are fat-soluble. Recently, it has become apparent that symbionts encode the majority of essential amino acid
phloem-feeding insects have independently evolved pathways (Fig. 2) and because plants are generally

© 2013 John Wiley & Sons Ltd


1478 A . K . H A N S E N and N . A . M O R A N

Gain of divergent nutritional essential amino acid co-symbiont amino acid


symbionts enabling sap feeding pathway gene loss complementation

gain of obligate symbiont A long-term obligate symbiont encodes gene gene lost
gain of obligate symbiont B
obligate co-symbiont encodes gene essential amino acid
gain of obligate co-symbiont C

Nutritional symbiont encoded factors

Influence insect nutrition Enable, maintain, and/or constrain host plant use

Insect encoded factors


host controlled amino acid host control of essential host control of symbiont
transport surrounding symbiont cells amino acid pathways cell density
outside

inside
amino acid/poly- amine/organocation host genes regulating the symbiont’s essential host killed symbiont cell
(APC) transporter amino acid pathways
live symbiont cell
obligate symbiont’s essential amino acid pathways
essential amino acid

Fig. 3 Mechanisms mediated by intracellular symbionts and insect hosts that potentially influence plant use by host insects. Proposed
mechanisms can either expand or contract insect host plant breadth. Upper panel: left, macroevolutionary patterns of gain and loss
of obligate symbionts; centre and right, complementarity of metabolic capabilities of cosymbionts. Lower panel: left, host transporter
structure generated by Pyre2 (Kelley & Sternberg 2009); centre, metabolism in the aphid bacteriocyte, redrawn from Hansen & Moran
(2011); right, host control of symbiont numbers, redrawn from Nishikori et al. (2009).

similar in amino acids available in phloem and xylem, lar gene content and order have been maintained
symbionts for the most part appear to have limited among strains, especially for essential amino acid bio-
influence on host plant breadth. In some insect species, synthesis, despite millions of years of divergence in
multiple strains of heritable symbionts have been separate insect lineages of insect hosts feeding on diver-
sequenced; these cases include Buchnera of aphids, gent host plants (Tamas 2002; Degnan et al. 2011; Lam-
Carsonella ruddii of psyllids and Sulcia muelleri of auc- elas et al. 2011; Sloan & Moran 2012a; McCutcheon &
henorrhynchan sap feeders. In each of these cases, simi- Moran 2012, van Ham et al. 2003).

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1479

Although there are broad similarities across plants in encoded in genomes of cosymbionts (Sloan & Moran
sap composition and across heritable symbionts in 2012a; Fig. 2). However, cosymbionts are lacking in
nutrient-provisioning capabilities, plant species do show some psyllid species, and how these psyllids acquire
some variation in sap composition, and several symbio- histidine and tryptophan is not clear (Nakabachi et al.
nts lack genes for multiple steps in amino acid biosyn- 2006; Sloan & Moran 2012a). Potentially some psyllids,
thesis. This variation raises the possibility that symbiont such as gall-forming species of Pachypsylla, obtain
biosynthetic capabilities partially constrain host plant enhanced nutrition through manipulation of the host
breadth. For several cases, genes that encode the major- plant. As mentioned above, a majority of sequenced
ity of the methionine, tryptophan, histidine or arginine genomes of symbionts of phloem-feeding insects also
pathways have been lost (Fig. 2). lack genes for arginine biosynthesis, specifically for the
In particular, pathways for the sulphur-containing intermediate metabolite ornithine; nevertheless, most
amino acids methionine and cysteine, and for sulphur retain genes encoding the enzymes that catalyse the
assimilation through sulphate reduction, are not reactions for arginine synthesis from ornithine (Fig. 2),
encoded in 70% of symbionts from phloem feeders suggesting that insect- and/or plant-derived ornithine
(Fig. 2) and are also absent from the small genome of is required for arginine biosynthesis.
the sole symbiont of a mesophyll-feeding scale insect In conclusion, obligate symbionts appear to lose some
(Sabree et al. 2013). These losses present a conundrum genes within essential amino acid pathways if these
because animals do not encode enzymes to reduce amino acids or their intermediates can be provisioned
sulphur for incorporation into the essential amino acids by the insect or cosymbiont or directly from the diet.
methionine and cysteine, in addition to other essential Loss of symbiont genes is irreversible, because long-
sulphur-containing metabolites. In contrast, most term symbiont genomes do not obtain genes through
xylem-feeding insects studied to date possess cos- horizontal transfer (Tamas 2002; Degnan et al. 2011;
ymbionts that retain most steps in the pathways for Lamelas et al. 2011). Ultimately, such gene loss in sym-
methionine biosynthesis and sulphur assimilation. bionts may constrain insects to a specific host plant that
Potentially, xylem sap lacks sufficient reduced sulphur is enriched in particular nutrients, unless these can be
or sulphur-containing organic molecules and thus obtained from cosymbionts and/or insect-encoded
imposes a more stringent requirement that obligate enzymes (Fig. 3). Potentially this can lead to increasing
symbionts assimilate inorganic sulphur. This difference numbers of obligate associations with symbionts that
between xylem and phloem feeders might reflect the provision different required nutrients, as seen in the
fact that a major form of reduced sulphur, S-methyl- cosymbionts of Auchenorrhyncha (McCutcheon &
methionine, is transported primarily in the phloem not Moran 2012). It is unclear whether symbiont gain or
xylem, especially when sulphur is reduced in mature replacements can increase host plant breadth.
leaves. But the site of sulphur reduction varies depend-
ing on environmental conditions and plant species
Dynamic potential of the integrated insect–microbe
(Bourgis et al. 1999; Tan et al. 2010). For example, in
amino acid metabolism
legumes such as pea (Pisum sativum), sulphur reduction
occurs in the roots, and S-methyl-methionine is then The insect’s own biosynthetic capabilities, and their
transported from xylem to phloem (Buchner et al. 2004). regulation, must be integrated with those of nutrient-
In sum, it is not yet clear why certain obligate symbio- provisioning symbionts. This integration of insect and
nts retain or lose pathways for methionine, sulphur symbiont metabolic capabilities could affect dietary
reduction and cysteine. We hypothesize that the reten- nutritional requirements and thus insect host plant
tion and loss of these pathways are associated with the range. For example, in one pea aphid–Buchnera lineage,
mode of reduced sulphur translocation in the plant and a frameshift mutation was detected in argC, one of the
with the insect host’s ability to assimilate this reduced five genes required for biosynthesis of ornithine, an
sulphur into methionine and cysteine. intermediate in arginine biosynthesis (Vogel & Moran
In addition to the widespread loss of the sulphate 2010). This aphid clone incurred fitness costs when fed
assimilation pathway from phloem feeders, several on artificial diet lacking arginine (Vogel & Moran 2010).
other amino acid biosynthetic pathways are occasionally This arginine dependence was attributed not only to
absent. For example, pathways for tryptophan, histidine Buchnera’s frameshift in argC, but also to insect host-
and arginine biosynthesis are absent from Carsonella, encoded factors based on genetic crosses (Vogel &
the nutritional symbiont of psyllids (Nakabachi et al. Moran 2010). This observation fits with the fact that
2006; Sloan & Moran 2012a; Fig. 2). In some psyllid ornithine may be either present in the phloem sap or
species, such as Heteropsylla cubana and Cternarytaina produced by the host insect, for example, through the
eucalypti, the tryptophan and/or arginine pathways are enzyme ornithine aminotransferase, which interconverts

© 2013 John Wiley & Sons Ltd


1480 A . K . H A N S E N and N . A . M O R A N

ornithine and glutamate and which is upregulated lated efficiently for host demands, this may limit the
in pea aphid bacteriocytes (Hansen & Moran 2011). insect’s survival and therefore persistence on particular
Therefore, in this example, the transport of ornithine host plants. Currently, only limited evidence, primarily
into Buchnera cells may be critical, and this transport is in aphids, has been found for an ability of obligate sym-
likely to be dependent on host-encoded products, as bionts to regulate essential amino acid pathways in
transporter genes are largely absent from Buchnera ge- response to insect nutritional demands (Moran et al.
nomes (Shigenobu et al. 2000) and are present in the 2003, 2005; Charles et al., 2006; Reymond et al., 2006).
aphid genome and upregulated in bacteriocytes (Han- Most symbiont genes regulating biosynthetic pathways
sen & Moran 2011; Poliakov et al. 2011; Price et al. have been lost in the course of genome reduction
2011). Thus, despite little variation in which amino acid (Shigenobu et al. 2000; McCutcheon & Moran 2012). The
pathways are encoded by symbionts, essential amino majority of genes encoding transcriptional regulators
acid requirements may vary due to variation in the for essential amino acid pathways are repressors, and
insect host genotype. For example, arginine dependence these are eliminated from symbiont genomes, possibly
may be impacted by variation among pea aphid geno- as an adaptation for overproducing the needed nutri-
types affecting capabilities for efficiently transporting ents (Moran et al. 2003). But metR, a transcription acti-
ornithine into Buchnera cells, whereas in aphid species vator, has been retained in genomes of some Buchnera
in which Buchnera has lost the ornithine pathway and of Baumannia. In the aphid Schizaphis graminum, the
(Fig. 4), the aphid may be able to efficiently transport Buchnera metR responds moderately to the demand for
ornithine to Buchnera. methionine resulting in increased transcription of metE,
Potentially, symbiont and/or insect regulation of encoding methionine synthase (Moran et al. 2003). In
amino acid biosynthetic pathways plays an important Buchnera, essential amino acid genes are located on the
role in host plant use. For example, if essential amino main chromosome, except for leuABCD and trpEG,
acid pathways can be regulated based on insect nutri- which are located on plasmids in some species (Shige-
tional demands, this may increase the suitability of nobu et al. 2000; Moran 2007). Interestingly, genes on
plants presenting different nutrient compositions in the leucine plasmid are regulated to enable synthesis of
different plant parts, times of year or plant taxa. Alter- leucine according to aphid demand (Vi~ nuelas et al.
natively, if certain essential amino acids are not regu- 2011). This study also found that leucine is a limiting

Ornithine-(Arg) Ile BCAA Phe AroAA


Suborder Superfamily Family Species Nutritional symbiont(s) argA, argAG argB argC argD argE ilvA ilvE aspC aroE
Auchenorrhynycha Cercopoidea Clastopteridae Clastoptera arizonana Sulcia and Zinderia

Cicadoidea Cicadidae Diceroprocta semicincta Sulcia and Hodgkinia

Membracoidea Cicadellidae Homalodisca vitripennis Sulcia and Baumannia

Sternorrhynycha Aphidoidea Aphididae Acyrthosiphon pisum Buchnera APS

Acyrthosiphon kondoi Buchnera AK

Uroleucon ambrosiae Buchnera Uamb

Schizaphis graminum Buchnera Sgram

Pemphigidae Baizongia pistaciae Buchnera Bp

Lachnidae Cinara cedri Buchnera Cc and Serratia symbiotica

Cinara tujafilina Buchnera Ct

Aleyrodoidea Aleyrodidae Bemisia tabaci Portiera

Psylloidea Aphalaridae Pachypsylla venusta Carsonella PV


Pachypsylla celtidis Carsonella PC
Ctenarytaina eucalypti Carsonella CE and SS
Ctenarytaina spatulata Carsonella CS
Psyllidae Heteropsylla texana Carsonella HT
Heteropsylla cubana Carsonella HC and SS

Coccoidea Pseudococcidae Planococcus citri Tremblaya and Moranella endobia

Heteroptera Pentatomoidea Plataspidae Megacopta punctatissima Ishikawaella

Fig. 4 Missing genes in essential amino acid pathways of intracellular nutritional symbionts. Missing ornithine biosynthesis genes
include argA, argAG, argB, argC, argD, argE. ilvA is an isoleucine biosynthesis gene. All branched-chain amino acid biosynthesis path-
ways include ilvE. aspC is a phenylalanine biosynthesis gene. Aromatic amino acid pathway genes include aroE. Blue, gene present
in symbiont; white, gene absent from symbiont and putative insect homolog may complement pathway.

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1481

factor for aphid growth and that higher leucine concen- threonine ammonia-lyase (IlvA, EC 4.3.1.19) was
trations increase the insect’s preference for a diet hypothesized to convert threonine into 2-oxobutanoate,
(Vi~nuelas et al. 2011). Plasmid-associated regulation of thus providing this metabolite to Buchnera as the sub-
essential amino acid production in obligate symbionts is strate for isoleucine biosynthesis. Second, the pea aphid
rare. Among systems for which genomes of symbionts homolog for a branched-chain amino acid transaminase
of sap-sucking insects are sequenced, only Buchnera and (IlvE, EC 2.6.1.42) was hypothesized to catalyse the final
Ishikawaella from the stink bug Megacopta punctatissima step of isoleucine, leucine and valine production. Third,
possess plasmids. the aphid homolog for aspartate transaminase (AspC,
Self-regulation of symbiont cell density may be EC 2.6.1.1) was predicted to catalyse the final step of
another mechanism that regulates essential amino acid phenylalanine biosynthesis (Wilson et al. 2010). Phenyl-
concentrations. In the unique nested symbioses of the alanine could then be converted into tyrosine via
mealybug Planococcus citri, the coresident symbiont another aphid enzyme (1.14.16.1). Transcriptome data
(Moranella) resides within cells of the nutritional symbi- from Hansen & Moran (2011) strongly supported all of
ont (Tremblaya). During different insect life stages, these predicted hypotheses as all three aphid homologs
Tremblaya was initially predicted to regulate Moranella’s were significantly upregulated in aphid bacteriocytes
cell densities (Kono et al. 2008). However, after both (Hansen & Moran 2011); however, ilvA was upregulated
symbiont genomes were sequenced, no genes were less than twofold (Hansen & Moran 2011), and an alter-
identified that would be capable of bacterial symbiont- native enzyme may produce 2-oxobutanoate (Poliakov
meditated regulation. Instead, the insect host may et al. 2011).
regulate symbiont cell numbers and organization Based on gene contents of symbionts, host and/or
(McCutcheon & von Dohlen 2011). In general, insect diet complementation appears widespread in phloem
hosts are hypothesized to regulate symbiont cell num- feeders in Sternorrhyncha and Heteroptera (Figs 3 and
bers, which fluctuate according to the insect life stages 4). Curiously, nutritional symbionts of xylem feeders
and morphs (Douglas & Dixon 1987; Braendle et al. from three separate superfamilies in Auchenorrhyncha
2003; Wolschin et al. 2004; Rio et al. 2006; Nishikori et al. do not appear to require insect genes to complement
2009; Stoll et al. 2010). any of their essential amino acid pathways (Fig. 4). As
The insect host may regulate symbiont-provisioned mentioned above, the first half of the arginine pathway
essential amino acid concentrations by regulating sym- is missing in 75% of phloem-feeding symbionts (12 of
biont cell density and/or by changing the activity of 16), yet genes are retained for the steps of this pathway
insect genes and transporters involved in the inte- subsequent to ornithine biosynthesis (Fig. 4). In con-
grated metabolism with the symbiont (Wilson et al. trast, the entire ornithine and arginine pathway is intact
2010; Hansen & Moran 2011; Price et al. 2011). If in nutritional symbionts of xylem-feeding insects.
insects can regulate symbiont production of essential Second, most phloem feeders (14 of 16) have lost the
amino acids in response to nutritional demands, then enzyme ilvA in the isoleucine pathway; however, this
the underlying insect genetic factors could be gained, gene is retained by symbionts of all xylem feeders
modified or lost, potentially influencing insect host (Fig. 4). Third, symbionts of all phloem feeders except
plant use (Fig. 3). Overall, we predict that the insect’s psyllids have lost ilvE, the final reaction synthesizing
genome and not that of the intracellular symbiont may each of the three branched-chain amino acids. Fourth,
be more important in influencing specific patterns of the terminal reaction of phenylalanine biosynthesis
host plant use. (aspC) converting phenylpyruvate to phenylalanine is
Recent evidence for the pea aphid–Buchnera system missing in all symbionts of phloem feeders but is pres-
suggests that insect regulation of the integrated metabo- ent in xylem-feeding symbionts (Fig. 4).
lism may be key in responding to variable amino acid This prevalence of insect host complementation in
availability in plant sap. More specifically, insect- phloem feeders may emerge because host genes are
encoded enzymes are hypothesized to collaborate with important in regulating arginine, phenylalanine and
the obligate symbiont in the biosynthesis of essential branched-chain amino acids primarily in phloem feed-
amino acids. For example, three enzymes integral to six ers and not xylem feeders. Conversely, insects within
essential amino acid pathways are missing from the Auchenorrhyncha may have lost host homologs for
genome of Buchnera of pea aphids (Shigenobu et al. complementing genes. Currently all Auchenorrhyncha
2000). These missing enzymes may be complemented for which symbiont sequences are available are xylem
by aphid-encoded homologs, based on aphid genomic feeders, and it will be of interest to see whether symbio-
data (Wilson et al. 2010), and supported by RNAseq nts of phloem-feeding Auchenorrhyncha show different
and proteomics data (Hansen & Moran 2011; Poliakov genetic repertoires. Only after more symbiont and insect
et al. 2011). For example, an aphid-encoded homolog of genomes have been sequenced from phloem- and

© 2013 John Wiley & Sons Ltd


1482 A . K . H A N S E N and N . A . M O R A N

xylem-feeding Auchenorrhyncha can this hypothesis be transmitted gut symbiont Ishikawaella capsulata has
adequately explored and tested. undergone genomic reduction comparable to that of
In summary, obligate intracellular symbionts of sap bacteriocyte-associated symbionts such as Buchnera in
feeders show little variation in amino acid biosynthetic aphids, and Ishikawaella retains a similar set of genes for
capabilities: they typically retain most essential amino the biosynthesis of essential amino acids, strongly sug-
acid pathways, and when particular genes are missing, gesting this function (Hosokawa et al. 2006; Nikoh et al.
the missing genes are present in genomes of a cosymbi- 2011). In addition, Ishikawaella contains a plasmid
ont and/or the insect host (Fig. 2). Only in the pea encoding a gene for oxalate decarboxylase, raising the
aphid has the insect host complementation of symbiont possibility that it provides the host with defence to the
provisioning been studied directly, so it is not yet clear widespread plant toxin oxalic acid.
whether there are general patterns with respect to the In some other herbivorous Heteroptera, gut symbio-
dynamics of essential amino acid regulation and trans- nts are acquired from the environment every genera-
port in nutrient-provisioning symbioses. Based on the tion, and this acquisition is sometimes highly selective
current information from complete genome sequences for a particular bacterial strain (Kikuchi et al. 2005,
of symbionts, insect enzymes appear to complement 2010, 2011; Hosokawa et al. 2012; Matsuura et al. 2012).
those of symbionts only in phloem feeders. Potentially, In one such case, the bean bug Riptortus pedestris
insect hosts are able to regulate symbiont provisioning (Heteroptera: Alydidae) is host to a specific Burkholderia
through control of nutrient transport and of symbiont strain (Kikuchi et al. 2005). When this insect occurs in
cell densities (Fig. 3). Although there is little empirical fields heavily treated with the insecticide fenitrothion,
evidence to date that heritable symbionts directly affect its Burkholderia symbiont shows capability to degrade
host plant specificity (see Box 1), the insect’s integrative fenitrothion, possibly due to gene uptake from nearby
metabolism in conjunction with nutritional symbiont(s) environmental bacteria, which show high levels of feni-
genotypes do have potential to influence host plant trothion degrading activity at such sites (Kikuchi et al.
utilization. 2012). This raises the possibility that a gut symbiont
with a dynamic genome is able to acquire novel capa-
bilities that enhance fitness of the host, a case that
Gut symbionts
parallels that of gut bacteria in humans acquiring genes
As in animals generally, almost all insects house for the degradation of specific foods consumed in spe-
microbes in the gut lumen. Most insect gut microbiomes cific populations (Hehemann et al. 2012).
are dominated by a small number of bacterial species, Bees feed exclusively on nectar, which is sugar rich,
with wide variation among species and higher taxa and and pollen, their sole source of protein. The gut micro-
few clear patterns based on surveys to date (Colman biota of adult worker honey bees Apis mellifera (Hyme-
et al. 2012). Much higher levels of community diversity noptera: Apidae) is somewhat unusual in being
have been recorded in some insect species, particularly transmitted through social contact within the hive and
species feeding on detritus or dead wood (e.g. Egert in consisting of a distinctive set of bacterial species spe-
et al. 2003; Warnecke et al. 2007; Andert et al. 2010; Reid cific to honey bees and to the related bumble bees
et al. 2011; Schauer et al. 2012). Although surveys of gut (Bombus species) (Martinson et al. 2012; Moran et al.
communities in herbivorous insects are still somewhat 2012). Based on a metagenomic study of this commu-
limited, a few species have been studied in depth. As nity, the genomes of these bacteria show evidence of
summarized below, rather few direct demonstrations of horizontal gene transfer and possess an excess of genes
symbiont roles in herbivory have been published to for metabolism of sugar and other carbohydrates (Engel
date. et al. 2012). One species, Gilliamella apicola (Kwong &
Among the simplest insect gut microbiomes are those Moran 2012), includes strains that encode pectate lyase
of species in a number of herbivorous families of and were demonstrated to be able to digest pectin in
Heteroptera, including Plataspidae, Pentatomidae, Bery- laboratory cultures (Engel et al. 2012). As pectin is a
tidae, Coreidae, Lygaeidae and Alydidae. Most species part of the cell wall that makes up one of the layers sur-
in these groups have guts with specialized midgut cae- rounding the protein-rich pollen germ, this pectinase
ca, which harbour a single bacterial type in each insect function may assist bees in obtaining protein from their
species (Fukatsu & Hosokawa 2002; Kikuchi et al. 2005, pollen diet. Interestingly, the pectate lyase-encoding
2010; Prado et al. 2009; Hosokawa et al. 2012; Matsuura genes in G. apicola are closely related to those from
et al. 2012). In some cases, these are transmitted mater- plant pathogens and appear to have been acquired
nally via symbiont capsules that are deposited with through horizontal gene transfer.
eggs and then ingested by new hatchlings. In the Although most bark beetles (Scolytidae) feed on tis-
plataspid bug Megacopta punctatissima, the maternally sues of dead plants, members of the genus Dendroctonus

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1483

Box 1
Is host plant specificity driven by symbionts in aphids?
The pea aphid is a species complex of recently diversified host races using different plant species in the legume
family (Ferrari et al. 2008). These races appear to be very recently established, possibly in association with host
plant shifts that coincided with post-Pleistocene warning and anthropogenic range expansion of several domesti-
cated legumes (Peccoud et al. 2009). Studies on sympatric populations of the races specialized to alfalfa (Medicago
sativa) and clover (Trifolium species) show that host plant specificity is determined primarily by host plant deter-
rents and assortative mating on food plants (Caillaud & Via 2000, 2012). Quantitative genetic studies, using genetic
crosses between the host plant races, showed that several aphid genes underlie differences in ability to use alfalfa
and clover (Hawthorne & Via 2001; Caillaud & Via 2012). These studies did not detect a maternal effect, as would
be expected if heritable symbionts had a substantial role in determining host plant use. Aphid genes appear to be
involved in differences among the other races of pea aphid as well: a recent genome-wide screen of divergent pea
aphid host races revealed 11 loci under divergent selection that are potentially responsible for differential host
plant usage by these two races; five of these loci may be involved in host plant adaptation and encode olfactory
and salivary proteins (Jaquiery et al. 2012).
It has been suggested that facultative symbionts of insect herbivores, particularly in the pea aphid system, may
help drive host plant adaptation resulting in specialized host-adapted races (Leonardo & Muiru 2003; Tsuchida
et al. 2004). A survey of facultative symbionts in pea aphid host races found significant associations between host
race and presence of particular symbionts (Ferrari et al. 2012). However, an association of a particular secondary
symbiont with a specific host plant race in the field does not imply that the symbiont is affecting host plant utiliza-
tion. Among the several alternative explanations are historical associations, because the secondary symbionts are
maternally transmitted and may happen to be coinherited with aphid genotypes able to use particular plants, and
also indirect effects of symbionts that impact host use. Indirect effects could include protection against natural ene-
mies (Oliver et al. 2003, 2005, 2009; Ferrari et al. 2004; Scarborough et al. 2005; Vorburger et al. 2010; Hansen et al.
2012) or thermal tolerance (Montllor et al. 2002); both are well documented and involve factors expected to differ
among plants. Direct experimental results regarding effects of facultative symbionts on plant host use are mixed.
One of the most prominent associations is between presence of the symbiont Regiella insecticola and use of clover as
a host plant (Ferrari et al. 2012). In one study, it was shown experimentally that R. insecticola increases pea aphid
performance on clover, but does not affect performance on fava bean (Vicia faba), a plant on which most pea aphid
host races perform well (Tsuchida et al. 2004). But in several other studies, using different clones of both aphid and
R. insecticola, this result has not been supported (Leonardo 2004; Ferrari et al. 2007; McLean et al. 2011).

attack living coniferous trees, feeding on the phloem biomass (Adams et al. 2013). Together these findings
tissue. Their food is poor in nitrogen and high in plant suggest that gut microbial associates in bark beetles
secondary metabolites, including abundant terpenes. probably contribute to nutrient provisioning, digestion
The microbial communities in the guts of Dendroctonus of plant polymers and detoxification of plant secondary
species contain a rather low diversity of bacteria (Mor- metabolites.
ales-Jimenez et al. 2009, 2012; Adams et al. 2013). Besides the relatively specialized cases of gut symbio-
Whole-animal and whole-gut preparations of these bee- ses of heteropterans, honey bees and Dendroctonus bark
tles show nitrogen fixation in laboratory assays, and iso- beetles, studies of gut microbiome composition in
lates of major members of this community are able to herbivorous insects are available for lepidopteran larvae
fix molecular nitrogen and to recycle uric acid waste and grasshoppers (Dillon & Charnley 2002; Sittenfeld
into usable amino acids (Morales-Jimenez et al. 2009, et al. 2002; Broderick et al. 2009; Robinson et al. 2010;
2012, 2013). In addition, isolates from the gut commu- Tang et al. 2012). Generally, these communities are
nity of Dendroctonus valens displayed cellulase activity highly variable even within a species, but they do con-
and thus potentially aid hosts in cellulose digestion tain somewhat distinctive sets of bacterial species,
(Morales-Jimenez et al. 2009). A metagenomic study of although relative abundances vary with the develop-
bacterial communities associated with Dendroctonus pon- mental stage and diet of the host (e.g. Broderick et al.
derosae, including both gut and environmental symbio- 2004; Robinson et al. 2010; Tang et al. 2012).
nts, revealed an elevated number of genes involved in A major challenge to insects that ingest whole plant
terpene degradation, relative to metagenomic data sets cells, including most foliage feeders with chewing
from other bacterial communities that degrade plant mouthparts, such as many lepidopteran larvae and bee-

© 2013 John Wiley & Sons Ltd


1484 A . K . H A N S E N and N . A . M O R A N

tles, is the digestion of plant cell walls. These contain Earlier studies in cigarette beetles, Lasioderma serricorne
polysaccharides such as cellulose, xylan, pectin and (Coleoptera: Anobiidae), suggested that gut fungi
mannan. The needed enzymatic activities were once neutralized toxins that were experimentally added to
considered to be generally lacking in animals, and genes laboratory diets. Specifically, some fungal isolates from
encoding such enzymes are not present in Drosophila, the guts were able to degrade the compounds, and
mosquitoes or Bombyx mori (Watanabe & Tokuda 2010). insect performance was decreased upon exposure to
However, some insects do possess endogenous genes fungicides and dietary toxins together (Dowd & Shen
that encode these enzymes: this was first discovered in 2011). Experiments aimed at testing a role of gut fungi
termites (Watanabe et al. 1998) and later found in other in breakdown of caffeine in the coffee berry borer
groups such as some beetles and crickets (Watanabe & Hypothenemus hampei (Coleoptera: Curculionidae) pro-
Tokuda 2010). The role of gut microbes in digestion is duced negative results, as the isolated fungi were inhib-
more intensively researched in insects such as termites ited themselves by caffeine (Vega et al. 2003). Earlier
and scarab beetles that feed on dead plant tissues (e.g. experiments on degradation of dietary terpenes in the
Warnecke et al. 2007; Hongoh et al. 2008; Andert et al. Douglas-fir tussock moth, Orgyia pseudotsugata, weighed
2010), but similar roles of gut microbes are plausible in against a regular role of gut bacteria (Andrews &
insects that feed on live plants. Some wood-feeding Spence 1980). A similar proposal has been made for
beetles (Cerambycidae) contain fungal gut symbionts gypsy moths: their gut communities sometimes contain
that have critical roles in digestion and that are some- Rhodococcus strains, some of which are known to break
times transmitted by maternal smearing of eggs down plant-derived monoterpene toxins (Broderick
(Gr€unwald et al. 2010); as these beetles are feeding on et al. 2004). However, direct demonstration that this
dead tissues, they are perhaps not technically herbi- activity occurs within the herbivore guts has not been
vores, but they can damage living plants. reported. Overall, a role of gut microbial species in
Some herbivorous insects also appear capable of pro- detoxification appears to have been proposed more
ducing cellulases and pectinases encoded in their own often than directly tested.
genomes; for example, the leaf beetle Phaedon cochleariae In summary, although a number of researchers have
(Chrysomelidae) possesses several enzymes capable of suggested that herbivorous insects might depend on
breakdown of plant polysaccharides (Kirsch et al. 2012). gut microorganisms for provisioning nutrients, diges-
In some cases, insect genes encoding these enzymes tion of plant tissues or neutralization of plant toxins,
may have been acquired through horizontal gene trans- surprisingly few clear cases have been reported. Possi-
fer from microbes, possibly from gut symbionts (See bly the lack of examples reflects a paucity of experimen-
Box 2); in other cases, they appear to be ancestral tal studies; however, roles of gut microbes in these
within large groups of insects. The current expansion in processes have been proposed for decades (Berenbaum
complete genome sequencing in diverse insect species 1988; Barbosa et al. 1991). More broadly, there is a very
will help to clarify which sets of these genes are ances- large literature on mechanisms by which herbivorous
tral in insects (and animals generally) and which are insects digest plant tissues or overcome plant chemical
acquired through horizontal gene transfer to specific defences. Possibly, gut microbes are simply not central
groups. players in many instances.
Gut microbes in herbivorous insects could also play a
role in detoxifying plant compounds. As for the case of
Environmental symbionts
digestive contributions, there is very little information
about this for herbivorous insects. Studies on velvet- Microbes living outside the insect’s body may directly
bean caterpillars, Anticarsia gemmatalis (Lepidoptera: influence insect–host plant utilization by providing the
Noctuidae) showed that survivorship and growth were insect with nutrients, modulating host plant recognition
improved when bacterial gut communities were intact, and response or detoxifying secondary plant com-
suggesting some beneficial role (Vis^ otto et al. 2009). pounds. A large diversity of microbes inhabit the leaf
These insects feed on leguminous plants that are rela- phyllosphere, and the taxonomic profiles of these
tively nitrogen rich but that produce protease inhibitors communities can be host plant-specific regardless of
able to prevent digestion of proteins. However, bacterial geographical location (Redford et al. 2010). Environmen-
species isolated from the A. gemmatalis gut secreted pro- tal bacteria can function as symbionts of herbivores and
teases that were relatively unaffected by the protease potentially contribute to insect–host plant use only if
inhibitors present in the plant tissues (Pilon et al. 2013). symbionts are transmitted by the insects themselves or
These observations fall short of proving that gut bacte- are persistently maintained in or around the insect’s
ria are crucial to overcoming these plant defences, but host plant tissue. We refer to these collectively as ‘envi-
they do provide suggestive evidence for such a role. ronmental symbionts’. Currently, only a few examples

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1485

Box 2
The role of horizontal gene transfer in enabling insect herbivory
Although it was once thought to be almost absent as a process affecting the evolution of animal genomes, acquisi-
tion of functional genes from foreign sources has now been found in a variety of cases (Moran & Jarvik 2010;
Acu~ na et al. 2012). A large proportion of these cases involved genes that might enable plant-feeding animals to be
more efficient as herbivores; for example, plant parasitic nematodes appear to have acquired vitamin biosynthesis
pathways and enzymes for cell wall degradation through incorporation of foreign genes into their own genomes
(Craig et al. 2008; Danchin et al. 2010; Paganini et al. 2012; Schuster & Sommer 2012). A well-supported case rele-
vant to the herbivorous lifestyle in insects is the acquisition of a mannanase-encoding gene by the coffee berry
borer (Coleoptera: Curculionidae) (Acu~ na et al. 2012). Mannan is the dominant polysaccharide in the coffee berries
that are the sole diet of these beetles, and the beetle genome has incorporated bacterial genes encoding mannanase
demonstrated to be functional in digestion of these molecules. Phylogenetic analysis indicated that the genes were
derived from a Bacillus species, potentially a bacterium in the gut or associated with the plant.
The complete genome sequence of the spider mite Tetranychus urticae contains several laterally transferred genes
that may facilitate feeding on the wide variety of plants used by this herbivorous arthropod (Grbic et al. 2011).
Among the genes for which the evidence of transfer was strong were genes encoding dioxygenases proposed to
function in breakdown of toxic plant aromatic compounds, cyanate lyase that might function to detoxify cyano-
genic glycosides produced by some plants, exo-fructosidases that are upregulated on certain host plants and cobal-
amin-independent methionine synthase, which could provide a source for this essential amino acid. Other genes
for which there is strong support for horizontal transfer to genomes of herbivorous arthropods include several
independent cases of carotenoid biosynthetic genes transferred from fungi: recipients include aphids that retain the
genes from their transfer to an aphid ancestor (Moran & Jarvik 2010), spider mites (Grbic et al. 2011; Altincicek
et al. 2012) and gall midges (Cobbs et al. 2013). Analysis of sequenced genomes for several lepidopteran species
(Bombyx mori, Heliconius melpomene and Danaus plexippus) revealed about 12 putative horizontal transfer events for
each genome, with transferred genes arriving from bacterial or fungal donors and at least some of the transfers
appearing to be ancient based on phylogenetic analyses of their distributions (Sun et al. 2013). Functions of genes
transferred to lepidopteran genomes mostly involve sugar, starch or amino acid metabolism and potential detoxifi-
cation enzymes (Li et al. 2011).
Numerous other studies using genomic or transcriptomic data have provided suggestive evidence for possible
microbe-to-insect horizontal transfers of genes encoding potential cell wall-degrading enzymes; the recipients have
been herbivorous insects ranging from beetles to moths to lygaeid bugs (e.g. Shen 2003; Kirsch et al. 2012; Sun et al.
2013). Some of these polysaccharide-degrading genes may result from relatively recent acquisition of microbial
genes facilitating herbivory, but others likely represent ancient and widespread animal genes that appear to be hor-
izontally transferred due to sampling artefacts and their absence from some of the initially sequenced insect ge-
nomes such as that of Drosophila (Pauchet et al. 2010; Calder on-Cortes et al. 2012). The extent to which the
acquisition of herbivorous lifestyles has depended on acquisition of foreign genes is thus still unclear.

of such associations have been described for herbivo- long-term obligate symbiosis feeding from specialized
rous insects; these primarily involve fungal symbionts hyphal tips. For example, complete genomes from two
that are reliably transmitted and subsequently localized leaf-cutter ant species, Acromyrmex echinatior and Atta
to insect galls, nests or galleries. cephalotes, reveal that the arginine biosynthesis pathway
In one such association, basidiomycete fungi and leaf- is not encoded; however, this pathway is encoded in
cutter ants (species in the genera Acromyrmex and Atta) other insects with complete metamorphosis, including
have an obligatory symbiosis together, in which ants other ant species (Nygaard et al. 2011; Suen et al. 2011).
culture the fungus for food in specialized gardens. These results suggest that leaf-cutter ant genomes may
Worker ants tend the fungus garden and provide fresh have lost genes encoding enzymes for arginine biosyn-
plant material for the fungal garden’s nutrients. In thesis, because the obligate fungal cultivars may
return, the symbiotic fungus provides the ant colony provide a reliable supply of this essential free amino
with its primary food source by producing specialized acid, analogous to aphids and their obligate symbiont,
hyphal tips (gongylidia) (H€ olldobler & Wilson 2010; Buchnera.
Aylward et al. 2012b; Mueller 2012). Leaf-cutter ants The symbiotic fungus garden converts plant biomass
may have specific nitrogen requirements because of this into nutrients for ants by breaking down a wide variety

© 2013 John Wiley & Sons Ltd


1486 A . K . H A N S E N and N . A . M O R A N

of plant polymers, potentially from a diverse array of fungi, which are transported tree to tree by their beetle
host plant polymer types (Gomes De Siqueira et al. vector and can grow within larval galleries. A variety
1998; D’Ettorre et al.2002; Richard et al. 2005; Silva of Dendroctonus bark beetle species that feed on fungus-
et al.2006a,b; Schiøtt et al. 2008; Erthal et al. 2009; De colonized phloem compared to noncolonized phloem
Fine Licht et al. 2010; Kooij et al. 2011). Some plant display higher larval survival, and their larval galleries
biomass degradation enzymes are upregulated in the are shorter, indicating that beetles do not have to
specialized gongylidia located in the middle of the gar- compensate for a low nutrient diet by consuming more
den, where ants feed. Ants then deposit their faecal plant material (Six 2012). These results indicate that
fluid containing active fungal-derived enzymes on top some bark beetles benefit from their stable fungal asso-
of the fresh leaf pulp substrate, aiding in leaf cell wall ciates, because fungal symbionts can provide necessary
digestion (Schiøtt et al. 2010; Moller et al. 2011). Plant nutrients, such as nitrogen and sterols, which are
digestive enzymes are also encoded and expressed in limited in phloem tissue (Six 2012).
bacterial microbes associated with fungal gardens (Suen Fungal associates of bark beetles benefit by colonizing
et al. 2010; Aylward et al. 2012a). However, it is an ephemeral resource early, such as phloem tissue,
currently unclear whether any of these bacterial taxa before competitive fungal saprophytes arrive (Paine
contribute nutrients to the ant. et al. 1997). Stable fungal associates of tree-killing bark
The live plant material leaf cutters collect may con- beetles generally do not overwhelm tree defences or
tain secondary metabolites that are toxic not only to the cause tree death for the insect host (Six & Wingfield
ant but also to their fungal cultivars (Berenbaum 1988). 2011); however, both the insect and fungus need to
As a result, leaf cutters select host plant leaves that are tolerate and/or detoxify their host plant’s allelochemi-
nontoxic towards themselves or their symbiotic fungi; cals to develop and compete with other fungal taxa on
nonhost plant leaves can have fungicidal activities phloem tissue (Paine et al. 1997). The genome sequence
towards the symbiotic fungi (Hubbell et al. 1983, 1984; of the beneficial fungal symbiont Grosmannia clavigera of
Howard 1988). In both the field and laboratory, high the mountain pine beetle (Dendroctonus ponderosae)
concentrations of tannins in leaves deter foraging leaf reveals that G. clavigera may detoxify a variety of host
cutters, and these concentrations inhibit fungal cultivar plant defences, including terpenoids using secondary
growth (Nichols-Orians 1991a,b). Nevertheless, the ant’s plant compounds as a carbon source (DiGuistini et al.
fungal cultivar can broaden the ant’s host plant breadth 2011). Also, as mentioned above in the section on gut
by producing enzymes that are important in detoxifying symbionts, a recent metagenomic study revealed that
plant secondary metabolites. For example, the fungal bacterial communities that associate with the same host
cultivar of leafcutters produces a specific laccase plants and phloem tissue as the bark beetle, D. pondero-
enzyme that is expressed highly in specialized gongyli- sae, encode genes for terpene and monoterpene degra-
dia that ants feed upon (De Fine Licht et al. 2013). Ants dation (Adams et al. 2013). Whether or not these
then ingest but do not digest this carbohydrate-active environmental microbes reliably help the bark beetle
enzyme, and the enzyme is then passed on through the detoxify host tree defences remains to be tested.
ant’s faecal fluid directly onto fresh leaf pulp on top of In another environmental fungal symbiont system,
the garden where it displays a high level of laccase some lineages of ambrosia gall midge (Diptera:
activity and oxidizes phenolics. Orthologous genes Cecidomyiidae, tribes Lasiopterini and Asphondyliini)
encoding this laccase enzyme from symbiotic and free- symbiotically associate with a monophyletic clade of
living fungi reveal that this enzyme has been positively fungi called Botryosphaeria dothidea (Janson et al. 2010).
selected in fungal cultivars that produce gongylidia, the This fungus species is also known as a widespread
common ancestor of higher attine ants and leaf cutters plant pathogen that infects a wide range of host plant
(De Fine Licht et al. 2013). It is assumed that this families (Heath & Stireman 2010; Janson et al. 2010).
enzyme has been modified for detoxification of phenols, Gall midge lineages that develop symbiotic associations
potentially leading to the major transition to active with B. dothidea deposit both fungal conidia and eggs
herbivory 8–12 million years ago in the leaf-cutter line- into host plant tissues, which subsequently develop into
age (De Fine Licht et al. 2013). galls that are internally lined with the symbiotic fungi.
Analogous to leaf-cutting ants, some fungal symbio- These fungal symbionts provide midges with nutrients
nts of bark beetles provide their hosts with nutrients such as sterols (Bissett & Borkent 1988; Rohfritsch 2008;
and potentially detoxify secondary compounds. Most Janson et al. 2009) and may provide protection against
bark beetles feed on dead or stressed trees; however, natural enemies (Weis 1982; Heath & Stireman 2010;
several species in the genus Dendroctonus can feed on Janson et al. 2010).
living or recently killed plant tissue (Raffa & Berryman A recent phylogenetic study reveals that gall midge
1983). Bark beetles are universally associated with lineages with symbiotic fungi have broader host plant

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1487

Box 3
Insect endosymbionts that influence host plant metabolism and defences
Microbes that associate with plants are generally viewed from the plant’s perspective and are classified as plant
pathogens, endophytes or mutualists, depending on the microbe’s relationship with the host plant. When microbial
associations are shared between an insect and plant partner, these categories become less defined and depend lar-
gely on the host organism of interest. A plant pathogen or endophyte that is vectored by an insect is generally
called a plant pathogen or endophyte respectively regardless of the impact on the insect host. But some of these
plant microbes closely associate with the insect in which they can be inherited and/or be regularly maintained,
and on which they have important effects. For example, when insects vector microbes to host plants, positive
effects on insect fitness tend to occur for efficient and highly specialized insect vector systems (reviewed in Wein-
traub & Beanland 2006; Stout et al. 2006; Kluth et al. 2002; Belliure et al. 2004). Some plant-associated microbes
(pathogenic or endophytic) can also be viewed as facultative symbionts of insect hosts, because some of these
microbes are vertically transmitted in insects, are highly host specific to a particular insect host and/or are persis-
tently maintained in insect populations (Hansen et al. 2008; Raddadi et al. 2011; Bressan et al. 2012; Caspi-Fluger
et al. 2012). Recently, a few studies have revealed that insect endosymbionts, including vectored plant microbes,
can influence plant signalling. For a more comprehensive review on insect–plant–microbe interactions, especially
for plant pathogens that are not vectored or intimately associated with insects, see Stout et al. (2006).
One insect–microbe system in which the microbe affects the insect–plant interface is the leafhopper Macrosteles
quadrilineatus, which vectors Aster Yellows phytoplasma, strain Witches’ Broom (AY-WB). When the leafhopper
feeds on AY-WB-infected host plants (Arabidopsis), the insect displays higher offspring production compared with
insects feeding on uninfected plants, primarily due to the phytoplasma’s modulation of plant defences (Sugio et al.
2011). More specifically, using transgenic lines of Arabidopsis, it was found that the phytoplasma’s effector protein
(SAP11) is responsible for destabilizing two Arabidopsis transcription factors, ultimately resulting in reduced expres-
sion of lipooxygenase 2 and decrease in subsequent jasmonic acid (JA) production, which is involved in the major
signalling pathway for upregulating defences against herbivores (Sugio et al. 2011).
In another specialized insect–plant pathogen system, the tomato psyllid Bactericera cockerelli vectors the alphapro-
teobacterium, Liberibacter psyllaurous, into its solanaceous host plants potato and tomato (Hansen et al. 2008). Liberib-
acter psyllaurous is associated with the plant disease called psyllid yellows but is also vertically transmitted at
moderate to high frequencies in psyllid eggs and maintained throughout psyllid development, increasing in titre as
the psyllid develops (Hansen et al. 2008; Casteel et al. 2012). When L. psyllaurous is inoculated into tomato alone
without the insect, tomato genes induced by jasmonic acid and salicylic acid signalling are suppressed, especially
transcripts of defence genes that are known to target insect pests (Casteel et al. 2012). Additionally, when both
psyllids and L. psyllaurous simultaneously infect tomato plants, dampened defensive host plant responses similar to
treatments of L. psyllaurous alone occur, especially for psyllid life stages that harbour the highest concentrations of
the bacterium. Collectively these results suggest that the bacterium may play a role in suppressing insect-related
host plant defences when L. psyllaurous is vectored into tomato by the psyllid. Other Liberibacter species may also
be important in psyllid–host plant interactions. For example, other Liberibacter species are vectored only by psyllids
into phloem tissue and are known as plant pathogens and endophytes (McClean & Oberholzer 1965; Halbert &
Manjunath 2004; Teixeira et al. 2005; Hansen et al. 2008; Raddadi et al. 2011).
The leaf-mining moth Phyllonorycter blancardella and its intracellular alphaproteobacterial endosymbiont Wolbachia
represent another insect–endosymbiont system that has been hypothesized to interact with the host plant; Wolbachia
is not known to infect plant tissue. The leaf-mining moth remains active in senescing apple leaves by feeding
within photosynthetically active green patches on the leaf. The green patches contain increased concentrations of
cytokinins and nutrients compared with senescing leaf margins (Giron et al. 2007). Cytokinins are plant hormones
involved in nutrient mobilization and inhibition of senescence and are thought to be involved in a variety of
insect–plant and microbe–plant interactions (Giron et al. 2007; Robert-Seilaniantz et al. 2007; Walters et al. 2008).
After treatment with antibiotics, the leaf miner cannot produce green patches in senescing leaves, and cytokinin
concentrations decline; in contrast, untreated leaf miners maintain Wolbachia infections, produce green patches and
display higher fitness (Kaiser et al. 2010). Therefore, symbiont presence in the leaf miner corresponds to green
patches and higher cytokinin concentrations in senescing apple leaves. In this system, it is unclear how Wolbachia
may be involved in green-patch production and if the moth, plant or symbiont produces this cytokinin (Kaiser
et al. 2010). Alternatively, Wolbachia may not be responsible; potentially the phenomenon is caused by a different
microbe present at a lower density and also eliminated during antibiotic treatments.

© 2013 John Wiley & Sons Ltd


1488 A . K . H A N S E N and N . A . M O R A N

Box 3 Continued
In a second lepidopteran system, when the western corn rootworm, Diabrotica virgifera, feeds on maize and is
infected with Wolbachia, plant defence-related genes are downregulated compared with plants fed on by antibiotic-
treated rootworms (Barr et al. 2010). As in the leaf-mining P. blancardella system above, it is unclear whether or
how symbionts directly influence the host plant, whether Wolbachia and not another symbiont is involved, and
whether microbes influence insect-related factors that in turn impact host plant defence signalling.
Finally, another indirect route by which insect endosymbionts may influence insect–host plant interactions is by
protecting insect hosts from natural enemies that would otherwise make a plant unavailable as a host. Protection
of herbivorous insects by endosymbionts has been well established, particularly for cases in aphids (reviewed in
Oliver et al. 2010). Such effects may be widespread for facultative symbionts, including symbionts from Enterobac-
teriaceae such as Hamiltonella defensa (Oliver et al. 2003), but also Wolbachia (Hedges et al. 2008; Teixeira et al. 2008)
and Spiroplasma species (Jaenike & Brekke 2011).

ranges than those of nonsymbiotic gall midges (Joy microbes in the environment to their food plants, as seen
2013). In addition, midge lineages with fungal symbio- in a few fungal symbionts. Otherwise, host plant-specific
nts are more species-rich at the tribe and genus levels microbes may be maintained more effectively by the
compared with related lineages lacking fungal symbio- host plant itself for its own benefit against natural
nts (Joy 2013). Collectively these results suggest that enemies. In some cases, however, plant microbes may
fungal symbionts may facilitate niche expansions and be plant pathogens that can influence insect host plant
diversification of their gall-forming insect hosts. But the use if they are reliably vectored by the insect and main-
underlying mechanism by which fungal symbionts tained as insect bacterial endosymbionts (See Box 3).
facilitate radiation of gall midges are not yet clear;
potentially multiple mechanisms are involved (e.g.
Concluding remarks
nutritional, behavioural preference, protection against
natural enemies and/or modulating plant defences). Insects have evolved many adaptations to overcome the
In addition to gall midges, fungal plant pathogens nutritional, digestive and defensive obstacles encoun-
that are also insect symbionts may facilitate host plant tered during herbivory. Major evolutionary transitions
exploitation, as suggested by studies of the grape berry towards insect herbivory potentially have also been
moth Lobesia botrana. In this system, the phytopathogen dependent on the aid of microbial symbionts. Microbes
Botrytis cinerea is vectored by the grape berry moth associate with all animals including insects, and they
(Fermaud & Le Menn 1992) and tunnelling moth larvae encode an enormous range of metabolic capabilities that
facilitate fungal penetration into grape berries (Fermaud dwarf and complement those of animal genomes. The
& Le Menn1989). Female grape berry moths preferen- ability of microbial symbionts to impact insect host
tially lay eggs on fungus-infected grape berries and dis- plant use has been proposed for decades, but experi-
play higher fitness on berries infected with fungi than mental evidence for most of these proposals has lagged
on uninfected berries (Mondy et al. 1998). Fitness bene- behind. Whether this is due to experimental and tech-
fits from fungi may be attributed to fungus-derived nological limitations, especially in the past, or to a
sterols provided to the insect (Mondy & Corio-Costet limited role of symbionts, is not yet clear.
2000). The possible roles played by symbionts in facilitating
In summary, many microbes persist in the herbivore’s herbivory largely depend on the location of the symbi-
environment including the phyllosphere and potentially ont relative to the insect body. The location, whether
interact with insects and plants. Experimental studies intracellular, in the gut or outside the body, heavily
demonstrating environmental symbionts of insects that influences the fidelity of transmission, persistence of the
are persistently maintained and that reliably influence association during development, insect host control and
insect–host plant use and evolution are extremely rare. microbe genome evolution, and these factors shape the
Perhaps environmental bacterial symbionts that are ben- potential contributions of symbionts. Intracellular sym-
eficial and influence insect–host plant use rapidly transi- bionts, especially of sap feeders, have been extensively
tion to an and insect location that can be maintained studied, and many of these symbiont genomes have
more efficiently such as the gut and/or inside the body been sequenced. Genome sequence analyses combined
and salivary glands. Complex insect behavioural and/or with experimental results show that intracellular symbi-
morphological adaptations may be required for insects onts contribute to insect nutrition by supplementing
to maintain transmission and localization of beneficial nutrients lacking in plant diets. Compared with most

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1489

Box 4
Outstanding questions regarding the role of symbiotic microorganisms in the use of plant tissues or plant
species by herbivorous insects
To what extent has horizontal gene transfer, from microbial sources to the insect genome, played a role in enabling
herbivorous lifestyles?
How frequently do gut microorganisms of herbivorous insects aid in digestion of plant tissues or in detoxification
of plant-derived compounds?
Are environmental symbionts (living outside the insect body) continuously maintained in most herbivorous insect
lineages or are they primarily associated with the host plant/phyllosphere? In either case, how often do they influence
host plant use?
To what extent are gut symbionts reliably transmitted between individuals of herbivorous insect species or reliably
acquired from environmental sources?
How often has the acquisition of a herbivorous lifestyle, or the ability to use particular plant tissues or species,
depended on the acquisition of foreign genes?
Overall, how often is adaptation for utilizing particular plants based on variation and innovation in genetic capabili-
ties of the insect versus variation in symbiont genotype and cosymbiont associations?
How common are vertically transmitted insect endosymbionts that play dual roles as plant pathogens/endophytes
in herbivorous insects?

gut symbionts and most environmental symbionts, rarely been found to affect host plant range. Perhaps
intracellular symbionts are relatively reliable in facilitat- transient, study system-specific effects of symbionts on
ing in host plant use over evolutionary time because insect–plant interactions will be documented; however,
they are heritable, and the host most likely controls symbiont-facilitated expansions of host plant range may
most metabolic interactions. Despite high fidelity be rare events.
in transmission, evolutionary novelty is limited in In order for a microbial symbiont to have a central
intracellular symbionts, which undergo extensive gene role in facilitating herbivory in general or the use
loss and do not acquire foreign genes. The gain of other of particular plant species in particular, it must be
symbionts and/or host regulation and control may consistently associated with an insect host. For micro-
enable hosts to maintain or expand abilities to feed on organisms in the gut and outside the insect body, the
host plants. Gut symbionts offer the potential advanta- consistency and persistence of associations of particular
ges of strain diversity and the dynamic genomes able to microbes with particular host lineages are still being
acquire new genes and functions, but they are probably explored and appear to vary enormously among insect
less reliably transmitted between hosts, except in cases species (e.g. Colman et al. 2012; Mueller 2012). More
with specific behavioural adaptations on the part of work is needed, and many of the outstanding
hosts. Their location could enable gut symbionts to play questions (Box 4) can now be addressed in studies
major roles in digestion of plant macromolecules and in exploiting molecular methodologies developed for
detoxification, but despite this potential, empirical stud- environmental microbiology, genomics and metage-
ies supporting these roles are limited. Environmental nomics.
symbionts have similar genome dynamics as gut symbi-
onts; however, their association with the insect and its
host plant may be less reliable unless complex insect
Acknowledgements
behaviour and/or morphological adaptations maintain The authors thank Gordon Bennett and Patrick Degnan
the symbiosis. Only a handful of insect systems are for helpful comments on the manuscript. We also thank
known in which environmental symbionts, primarily three anonymous reviewers for their helpful comments
fungi, aid in host plant use by supplementing nutrients and suggestions. Funding was provided by the US
and/or breaking down large plant polysaccharides or Department of Agriculture [2011-67012-30707] to A.H
toxins. and by Yale University.
Although a general role of symbionts in enabling
herbivory is clear, especially for sap feeders, there is
limited experimental evidence for an effect of symbionts
References
on specificity of insect host plant use. In particular, gut Acu~
na R, Padilla BE, Florez-Ramos CP et al. (2012) Adaptive
symbionts, although present in almost all insects, have horizontal transfer of a bacterial gene to an invasive insect

© 2013 John Wiley & Sons Ltd


1490 A . K . H A N S E N and N . A . M O R A N

pest of coffee. Proceedings of the National Academy of Sciences, Braendle C, Miura T, Bickel R et al. (2003) Developmental ori-
USA, 109, 4197–4202. gin and evolution of bacteriocytes in the aphid-Buchnera
Adams AS, Aylward FO, Adams SM et al. (2013) Mountain symbiosis. PLoS Biology, 1, E21.
pine beetles colonizing historical and na€ıve host trees are Bressan A, Terlizzi F, Credi R (2012) Independent origins of
associated with a bacterial community highly enriched in vectored plant pathogenic bacteria from arthropod-associated
genes contributing to terpene metabolism. Applied and Envi- Arsenophonus endosymbionts. Microbial Ecology, 63, 628–638.
ronmental Microbiology, 61, 759–768. Brodbeck BV, Mizell RF III, French WJ, Andersen PC, Aldrich
Akman G€ und€uz E, Douglas AE (2009) Symbiotic bacteria JH (1990) Amino acids as determinants of host preference
enable insect to use a nutritionally inadequate diet. Proceed- for the xylem feeding leafhopper, Homalodisca coagulata
ings of the Royal Society B: Biological Sciences, 276, 987–991. (Homoptera: Cicadellidae). Oecologia, 83, 338–345.
Altincicek B, Kovacs JL, Gerardo NM (2012) Horizontally trans- Broderick NA, Raffa KF, Goodman RM, Handelsman J (2004)
ferred fungal carotenoid genes in the two-spotted spider Census of the bacterial community of the gypsy moth larval
mite Tetranychus urticae. Biology Letters, 8, 253–257. midgut by using culturing and culture-independent meth-
Andersen PC, Brodbeck BV, Mizell RF (1989) Metabolism of ods. Applied and Environmental Microbiology, 70, 293–300.
amino acids, organic acids and sugars extracted from the Broderick NA, Robinson CJ, McMahon MD et al. (2009) Contri-
xylem fluid of four host plants by adult Homalodisca coagula- butions of gut bacteria to Bacillus thuringiensis-induced mor-
ta. Entomologia Experimentalis et Applicata, 503, 149–159. tality vary across a range of Lepidoptera. BMC Biology, 7, 11.
Andersen PC, Brodbeck BV, Mizell RF III (1992) Feeding by Buchner P (1965) Endosymbiosis of Animals with Plant Microor-
the leafhopper, Homalodisca coagulata, in relation to xylem ganisms. Interscience Publishers/John Wiley, New York, NY.
fluid chemistry and tension. Journal of Insect Physiology, Buchner P, Takahashi H, Hawkesford MJ (2004) Plant sulphate
38, 611–622. transporters: co-ordination of uptake, intracellular and long-
Andert J, Marten A, Brandl R, Brune A (2010) Inter- and intra- distance transport. Journal of Experimental Botany, 55, 1765–1773.
specific comparison of the bacterial assemblages in the hind- Caillaud MC, Via S (2000) Specialized feeding behavior influ-
gut of humivorous scarab beetle larvae (Pachnoda spp.). ences both ecological specialization and assortative mating in
FEMS Microbiology Ecology, 74, 439–449. sympatric host races of pea aphids. The American Naturalist,
Andrews RE, Spence KD (1980) Action of douglas fir tussock 156, 606–621.
moth larvae and their microflora on dietary terpenes. Applied Caillaud MC, Via S (2012) Quantitative genetics of feeding
and Environmental Microbiology, 40, 959–963. behavior in two ecological races of the pea aphid, Acyrthosi-
Aylward FO, Burnum KE, Scott JJ et al. (2012a) Metagenomic phon pisum. Heredity, 108, 211–218.
and metaproteomic insights into bacterial communities Calderon-Cortes N, Quesada M, Watanabe H, Cano-Camacho
in leaf-cutter ant fungus gardens. The ISME Journal, 6, H, Oyama K (2012) Endogenous plant cell wall digestion: a
1688–1701. key mechanism in insect evolution. Annual Review of Ecology,
Aylward FO, Currie CR, Suen G (2012b) The evolutionary Evolution, and Systematics, 43, 45–71.
innovation of nutritional symbioses in leaf-cutter ants. Caspi-Fluger A, Inbar M, Mozes-Daube N et al. (2012) Horizon-
Insects, 3, 41–61. tal transmission of the insect symbiont Rickettsia is plant-
Barbosa P, Krischik VA, Jones CG (1991) Microbial Mediation of mediated. Proceedings of the Royal Society B: Biological Sciences,
Plant-Herbivore Interactions. Wiley-Interscience, New York, 279, 1791–1796.
NY. Casteel CL, Hansen AK, Walling LL, Paine TD (2012) Manipu-
Barr KL, Hearne LB, Briesacher S, Clark TL, Davis GE (2010) lation of plant defense responses by the tomato psyllid
Microbial symbionts in insects influence down-regulation of (Bactericerca cockerelli) and its associated endosymbiont Can-
defense genes in maize. PLoS One, 5, e11339. didatus Liberibacter psyllaurous. PLoS One, 7, e35191.
Baumann P (2005) Biology of bacteriocyte-associated endos- Charles H, Calevro F, Viñuelas J, Fayard J-M, Rahbé Y (2006)
ymbionts of plant sap-sucking insects. Annual Review of Codon usage bias and tRNA over-expression in Buchnera
Microbiology, 59, 155–189. aphidicola after aromatic amino acid nutritional stress on its
Belliure B, Janssen A, Maris PC, Peters D, Sabelis MW (2004) host Acyrthosiphon pisum. Nucleic Acids Research, 34, 4583–
Herbivore arthropods benefit from vectoring plant viruses. 4592.
Ecology Letters, 8, 70–79. Clark EL, Karley AJ, Hubbard SF (2010) Insect endosymbionts:
Berenbaum MR (1988) Allelochemicals in insect-microbe-plant manipulators of insect herbivore trophic interactions? Pro-
Interactions; agents provocateurs in the coevolutionary arms toplasma, 244, 25–51.
race. In: Novel Aspects of Insect-Plant Interactions (eds Barbosa Cobbs C, Heath J, Stireman JO, Abbot P (2013) Carotenoids in
P, Letourneau DK), pp. 97–123. Wiley-Interscience, New unexpected places: Gall midges, lateral gene transfer and
York, NY. carotenoid biosynthesis in animals. Molecular Phylogenetics
Bernays EA, Chapman RF (1994) Host Plant Selection by and Evolution, 68, 221–228.
Phytophagous Insects (Contemporary Topics in Entomology). Colman DR, Toolson EC, Takacs-Vesbach CD (2012) Do diet
Springer, New York, NY. and taxonomy influence insect gut bacterial communities?
Bissett J, Borkent A (1988) Coevolution of Fungi with Plants and Molecular Ecology, 21, 5124–5137.
Animals. Academic Press, New York, NY. Conord C, Despres L, Vallier A et al. (2008) Long-term evolu-
Bourgis F, Roje S, Nuccio ML et al. (1999) S-methylmethionine tionary stability of bacterial endosymbiosis in Curculionoi-
plays a major role in phloem sulfur transport and is synthe- dea: additional evidence of symbiont replacement in the
sized by a novel type of methyltransferase. The Plant Cell, 11, Dryophthoridae family. Molecular Biology and Evolution, 25,
1485–1497. 859–868.

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1491

Craig JP, Bekal S, Hudson M et al. (2008) Analysis of a horizon- Scarabaeidae). Applied and Environmental Microbiology, 69,
tally transferred pathway involved in vitamin B6 biosynthe- 6659–6668.
sis from the soybean cyst nematode Heterodera glycines. Engel P, Martinson VG, Moran NA (2012) Functional diversity
Molecular Biology and Evolution, 25, 2085–2098. within the simple gut microbiota of the honey bee. Proceedings
Cryan JR, Urban JM (2012) Higher-level phylogeny of the of the National Academy of Sciences, USA, 109, 11002–11007.
insect order Hemiptera: is Auchenorrhyncha really paraphy- Erthal M, Silva CP, Cooper RM, Samuels RI (2009) Hydrolytic
letic? Systematic Entomology, 37, 7–21. enzymes of leaf-cutting ant fungi. Comparative Biochemistry
Danchin EGJ, Rosso M-N, Vieira P et al. (2010) Multiple lateral and Physiology. Part B, Biochemistry & Molecular Biology, 152,
gene transfers and duplications have promoted plant parasit- 54–59.
ism ability in nematodes. Proceedings of the National Academy Farrell B (1998) “Inordinate Fondness” explained: why are
of Sciences, USA, 107, 17651–17656. there so many beetles? Science, 281, 555–559.
De Fine Licht HH, Schiott M, Mueller UG, Boomsa JJ (2010) Febvay GR, Liadouze I, Guillaud J, Bonnot G (1995) Analysis
Evolutionary transitions in enzyme activity of ant fungus of energetic amino acid metabolism in Acyrthosiphon pisum: a
gardens. Evolution, 64, 2055–2069. multidimensional approach to amino acid metabolism in
De Fine Licht HH, Schiøtt M, Rogowska-Wrzesinska A et al. aphids. Archives of Insect Biochemistry and Physiology, 29,
(2013) Laccase detoxification mediates the nutritional alliance 45–69.
between leaf-cutting ants and fungus-garden symbionts. Pro- Feldhaar H (2011) Bacterial symbionts as mediators of ecologi-
ceedings of the National Academy of Sciences, USA, 110, 583– cally important traits of insect hosts. Ecological Entomology,
587. 36, 533–543.
Degnan PH, Ochman H, Moran NA (2011) Sequence conserva- Felton GW, Tumlinson JH (2008) Plant–insect dialogs: complex
tion and functional constraint on intergenic spacers in interactions at the plant–insect interface. Current Opinion in
reduced genomes of the obligate symbiont Buchnera. PLoS Plant Biology, 11, 457–463.
Genetics, 7, e1002252. Fermaud M, Le Menn R (1989) Association of Botrytis cinerea
D’Ettorre P, Mora P, Dibangou V, Rouland C, Errard C (2002) with grape berry moth larvae. Phytopathology, 79, 651–656.
The role of the symbiotic fungus in the digestive metabolism Fermaud M, Le Menn R (1992) Transmission of Botrytis cinerea
of two species of fungus-growing ants. Journal of Comparative to grapes by grape berry moth larvae. Phytopathology, 82,
Physiology B, Biochemical, Systemic, and Environmental Physiol- 1393–1398.
ogy, 172, 169–176. Ferrari J, Darby AC, Daniell TJ, Godfray HCJ, Douglas AE
DiGuistini S, Wang Y, Liao NY et al. (2011) Genome and tran- (2004) Linking the bacterial community in pea aphids with
scriptome analyses of the mountain pine beetle-fungal sym- host-plant use and natural enemy resistance. Ecological Ento-
biont Grosmannia clavigera, a lodgepole pine pathogen. mology, 29, 60–65.
Proceedings of the National Academy of Sciences, USA, 108, Ferrari J, Scarborough CL, Godfray HCJ (2007) Genetic varia-
2504–2509. tion in the effect of a facultative symbiont on host-plant use
Dillon R, Charnley K (2002) Mutualism between the desert by pea aphids. Oecologia, 153, 323–329.
locust Schistocerca gregaria and its gut microbiota. Research in Ferrari J, Via S, Godfray HCJ (2008) Population differentiation
Microbiology, 153, 503–509. and genetic variation in performance on eight hosts in the
Douglas AE (1988) Sulphate utilization in an aphid symbiosis. pea aphid complex. Evolution, 62, 2508–2524.
Insect Biochemistry, 18, 599–605. Ferrari J, West JA, Via S, Godfray HCJ (2012) Population
Douglas AE (1998) Nutritional interactions in insect-microbial genetic structure and secondary symbionts in host-associated
symbioses: aphids and their symbiotic bacteria Buchnera. populations of the pea aphid complex. Evolution, 66, 375–390.
Annual Review of Entomology, 43, 17–37. Frago E, Dicke M, Godfray HCJ (2012) Insect symbionts as
Douglas AE, Dixon AFG (1987) The mycetocyte symbiosis of hidden players in insect–plant interactions. Trends in Ecology
aphids: variation with age and morph in virginoparae of and Evolution, 27, 705–711.
Megoura viciae and Acyrthosiphon pisum. Journal of Insect Physi- Fukatsu T, Hosokawa T (2002) Capsule-transmitted gut symbi-
ology, 33, 109–113. otic bacterium of the Japanese common plataspid stinkbug,
Douglas AE, Prosser WA (1992) Synthesis of the essential Megacopta punctatissima. Applied and Environmental Microbiol-
amino acid tryptophan in the pea aphid (Acyrthosiphon pi- ogy, 68, 389–396.
sum) symbiosis. Journal of Insect Physiology, 38, 565–568. Futuyma DJ, Agrawal AA (2009) Evolutionary history and spe-
Douglas AE, Bouvaine S, Russell RR (2011) How the insect cies interactions. Proceedings of the National Academy of
immune system interacts with an obligate symbiotic bacte- Sciences, USA, 106, 18043–18044.
rium. Proceedings of the Royal Society B: Biological Sciences, 278, Giron D, Kaiser W, Imbault N, Casas J (2007) Cytokinin-medi-
333–338. ated leaf manipulation by a leafminer caterpillar. Biology
Dowd PF, Shen SK (2011) The contribution of symbiotic yeast Letters, 3, 340–343.
to toxin resistance of the cigarette beetle (Lasioderma serricor- Gomes De Siqueira C, Bacci M, Pagnocca F, Bueno O, Hebling
ne). Entomologia Experimentalis et Applicata, 56, 241–248. M (1998) Metabolism of plant polysaccharides by Leucoagari-
Downie DA, Gullan PJ (2005) Phylogenetic congruence of cus gongylophorus, the symbiotic fungus of the leaf-cutting
mealybugs and their primary endosymbionts. Journal of Evo- ant Atta sexdens L. Applied and Environmental Microbiology, 64,
lutionary Biology, 18, 315–324. 4820–4822.
Egert M, Wagner B, Lemke T, Brune A, Friedrich MW (2003) Gosalbes MJ, Lamelas A, Moya A, Latorre A (2008) The strik-
Microbial community structure in midgut and hindgut of the ing case of tryptophan provision in the cedar aphid Cinara
humus-feeding larva of Pachnoda ephippiata (Coleoptera: cedri. Journal of Bacteriology, 190, 6026–6029.

© 2013 John Wiley & Sons Ltd


1492 A . K . H A N S E N and N . A . M O R A N

Grbic M, Van Leeuwen T, Clark RM et al. (2011) The genome Jaenike J, Brekke TD (2011) Defensive endosymbionts: a cryptic
of Tetranychus urticae reveals herbivorous pest adaptations. trophic level in community ecology. Ecology Letters, 14,
Nature, 479, 487–492. 150–155.
Gr€unwald S, Pilhofer M, H€ oll W (2010) Microbial associations Janson EM, Stireman JO, Singer MS, Abbot P (2008) Phytopha-
in gut systems of wood- and bark-inhabiting longhorned gous insect-microbe mutualisms and adaptive evolutionary
beetles [Coleoptera: Cerambycidae]. Systematic and Applied diversification. Evolution, 62, 997–1012.
Microbiology, 33, 25–34. Janson EM, Grebenok RJ, Behmer ST, Abbot P (2009) Same
Halbert SE, Manjunath KL (2004) Asian citrus psyllids (Stern- host-plant, different sterols: variation in sterol metabolism in
orrhyncha: Psyllidae) and greening disease of citrus: a litera- an insect herbivore community. Journal of Chemical Ecology,
ture review and assessment of risk in Florida. The Florida 35, 1309–1319.
Entomologist, 3, 330–353. Janson EM, Peeden ER, Stireman JO III, Abbot P (2010) Symbi-
van Ham RCHJ, Kamerbeek J, Palacios C et al. (2003) Reduc- ont-mediated phenotypic variation without co-evolution in
tive genome evolution in Buchnera aphidicola. Proceedings of an insect–fungus association. Journal of Evolutionary Biology,
the National Academy of Sciences, USA, 100, 581–586. 23, 2212–2228.
Hansen AK, Moran NA (2011) Aphid genome expression Jaquiery J, Stoeckel S, Nouhaud P et al. (2012) Genome scans
reveals host-symbiont cooperation in the production of reveal candidate regions involved in the adaptation to host plant
amino acids. Proceedings of the National Academy of Sciences, in the pea aphid complex. Molecular Ecology, 21, 5251–5264.
USA, 108, 2849–2854. Joy JB (2013) Symbiosis catalyses niche expansion and diversifi-
Hansen AK, Trumble JT, Stouthamer R, Paine TD (2008) A cation. Proceedings of the Royal Society B: Biological Sciences,
new Huanglongbing Species, “Candidatus Liberibacter psyl- 280, 2212–2228.
laurous”, found to infect tomato and potato, is vectored by Kaiser W, Huguet E, Casas J, Commin C, Giron D (2010) Plant
the psyllid Bactericera cockerelli (Sulc). Applied and Environ- green-island phenotype induced by leaf-miners is mediated
mental Microbiology, 74, 5862–5865. by bacterial symbionts. Proceedings of the Royal Society B:
Hansen AK, Vorburger C, Moran NA (2012) Genomic basis of Biological Sciences, 277, 2311–2319.
endosymbiont-conferred protection against an insect parasit- Kelley LA, Sternberg MJE (2009) Protein structure prediction
oid. Genome Research, 22, 106–114. on the web: a case study using the Phyre server. Nature
Hawthorne DJ, Via S (2001) Genetic linkage of ecological spe- Protocols, 4, 363–371.
cialization and reproductive isolation in pea aphids. Nature, Kikuchi Y, Meng X-Y, Fukatsu T (2005) Gut symbiotic bacte-
412, 904–907. ria of the genus Burkholderia in the broad-headed bugs
Heath JJ, Stireman JO III (2010) Dissecting the association Riptortus clavatus and Leptocorisa chinensis (Heteroptera:
between a gall midge, Asteromyia carbonifera, and its symbi- Alydidae). Applied and Environmental Microbiology, 71,
otic fungus, Botryosphaeria dothidea. Entomologia Experimentalis 4035–4043.
et Applicata, 137, 36–49. Kikuchi Y, Hosokawa T, Fukatsu T (2010) An ancient but pro-
Hedges LM, Brownlie JC, O’Neill SL, Johnson KN (2008) Wol- miscuous host–symbiont association between Burkholderia
bachia and virus protection in insects. Science, 322, 702. gut symbionts and their heteropteran hosts. The ISME Jour-
Hehemann J-H, Kelly AG, Pudlo NA, Martens EC, Boraston nal, 5, 446–460.
AB (2012) Bacteria of the human gut microbiome catabolize Kikuchi Y, Hosokawa T, Fukatsu T (2011) Specific devel-
red seaweed glycans with carbohydrate-active enzyme opmental window for establishment of an insect-microbe
updates from extrinsic microbes. Proceedings of the National gut symbiosis. Applied and Environmental Microbiology, 77,
Academy of Sciences, USA, 109, 19786–19791. 4075–4081.
H€olldobler B, Wilson EO (2010) The Leafcutter Ants: Civilization Kikuchi Y, Hayatsu M, Hosokawa T et al. (2012) Symbiont-
by Instinct. W. W. Norton, New York, NY. mediated insecticide resistance. Proceedings of the National
Hongoh Y, Sharma VK, Prakash T et al. (2008) Genome of an Academy of Sciences, USA, 109, 8618–8622.
endosymbiont coupling N2 fixation to cellulolysis within Kirsch R, Wielsch N, Vogel H et al. (2012) Combining proteo-
protist cells in termite gut. Science, 322, 1108–1109. mics and transcriptome sequencing to identify active plant-
Hosokawa T, Kikuchi Y, Nikoh N, Shimada M, Fukatsu T cell-wall-degrading enzymes in a leaf beetle. BMC Genomics,
(2006) Strict host-symbiont cospeciation and reductive gen- 13, 587.
ome evolution in insect gut bacteria. PLoS Biology, 4, Kluth S, Kruess A, Tscharntke T (2002) Insects as vectors of
e337. plant pathogens: mutualistic and antagonistic interactions.
Hosokawa T, Kikuchi Y, Nikoh N, Fukatsu T (2012) Polyphyly Oecologia, 133, 193–199.
of gut symbionts in stinkbugs of the family Cydnidae. Kölsch G, Pedersen BV (2010) Can the tight co-speciation
Applied and Environmental Microbiology, 78, 4758–4761. between reed beetles (Col., Chrysomelidae, Donaciinae) and
Howard JJ (1988) Leafcutting and diet selection: relative influ- their bacterial endosymbionts, which provide cocoon mate-
ence of leaf chemistry and physical features. Ecology, 69, rial, clarify the deeper phylogeny of the hosts? Molecular
250–260. Phylogenetics and Evolution, 54, 810–821.
Hubbell SP, Wiemer DF, Adejare A (1983) An antifungal terpe- Kono M, Koga R, Shimada M, Fukatsu T (2008) Infection
noid defends a neotropical tree (Hymenaea) against attack by dynamics of coexisting Beta- and Gammaproteobacteria in
fungus-growing ants (Atta). Oecologia, 60, 321–327. the nested endosymbiotic system of mealybugs. Applied and
Hubbell SP, Howard JJ, Wiemer DF (1984) Chemical leaf repel- Environmental Microbiology, 74, 4175–4184.
lency to an attine ant: seasonal distribution among potential Kooij PW, Schiott M, Boomsma JJ, De Fine Licht HH (2011)
host plant species. Ecology, 65, 1067–1076. Rapid shifts in Atta cephalotes fungus-garden enzyme activity

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1493

after a change in fungal substrate (Attini, Formicidae). Insec- McCutcheon JP, McDonald BR, Moran NA (2009) Convergent
tes Sociaux, 58, 145–151. evolution of metabolic roles in bacterial co-symbionts of
Kwong WK, Moran NA (2012) Cultivation and characterization insects. Proceedings of the National Academy of Sciences, USA,
of the gut symbionts of honey bees and bumble bees: Snodg- 106, 15394–15399.
rassella alvi gen. nov., sp. nov., a member of the Neisseriaceae McFall-Ngai M, Hadfield MG, Bosch TCG et al. (2013) Animals
family of the Betaproteobacteria; and Gilliamella apicola gen. in a bacterial world, a new imperative for the life sciences.
nov., sp. nov., a member of Orbaceae fam. nov., Orbales ord. Proceedings of the National Academy of Sciences, USA, 110,
nov., a sister taxon to the Enterobacteriales order of the 3229–3236.
Gammaproteobacteria. International Journal of Systematic and McLean AHC, van Asch M, Ferrari J, Godfray HCJ (2011)
Evolutionary Microbiology, 63, 2008–2018. Effects of bacterial secondary symbionts on host plant use in
Labandeira CC, Sepkoski JJ (1993) Insect diversity in the fossil pea aphids. Proceedings of the Royal Society B: Biological Sci-
record. Science, 261, 310–315. ences, 278, 760–766.
Lamelas A, Gosalbes MJ, Moya A, Latorre A (2011) New Mitter C, Farrell B, Wiegmann B (1988) The phylogenetic study
clues about the evolutionary history of metabolic losses in of adaptive zones: has phytophagy promoted insect diversifi-
bacterial endosymbionts, provided by the genome of Buch- cation? The American Naturalist, 132, 107–128.
nera aphidicola from the aphid Cinara tujafilina. Applied and Mittler TE (1971) Dietary amino acid requirements of the aphid
Environmental Microbiology, 77, 4446–4454. Myzus persicae affected by antibiotic uptake. The Journal of
Leonardo TE (2004) Removal of a specialization-associated Nutrition, 101, 1023–1028.
symbiont does not affect aphid fitness. Ecology Letters, 7, Moller IE, De Fine Licht HH, Harholt J, Willats WGT, Boom-
461–468. sma JJ (2011) The dynamics of plant cell-wall polysaccharide
Leonardo TE, Muiru GT (2003) Facultative symbionts are decomposition in leaf-cutting ant fungus gardens. PLoS One,
associated with host plant specialization in pea aphid 6, e17506.
populations. Proceedings of the Royal Society B: Biological Mondy N, Corio-Costet M-F (2000) The response of the grape
Sciences, 270, S209–S212. berry moth (Lobesia botrana) to a dietary phytopathogenic
Li Z-W, Shen Y-H, Xiang Z-H, Zhang Z (2011) Pathogen-origin fungus (Botrytis cinerea): the significance of fungus sterols.
horizontally transferred genes contribute to the evolution of Journal of Insect Physiology, 46, 1557–1564.
Lepidopteran insects. BMC Evolutionary Biology, 11, 356. Mondy N, Charrier B, Fermaud M, Pracros P, Corio-Costet M-F
Macdonald SJ, Thomas GH, Douglas AE (2011) Genetic and (1998) Mutualism between a phytopathogenic fungus (Botry-
metabolic determinants of nutritional phenotype in an tis cinerea) and a vineyard pest (Lobesia botrana). Positive
insect-bacterial symbiosis. Molecular Ecology, 20, 2073–2084. effects on insect development and oviposition behaviour.
Macdonald SJ, Lin GG, Russell CW, Thomas GH, Douglas AE Comptes Rendus de l’Academie des Sciences - Series III - Sciences
(2012) The central role of the host cell in symbiotic nitrogen de la Vie, 321, 665–671.
metabolism. Proceedings of the Royal Society B: Biological Montllor CB, Maxmen A, Purcell AH (2002) Facultative bacte-
Sciences, 279, 2965–2973. rial endosymbionts benefit pea aphids Acyrthosiphon pisum
Martinson VG, Moy J, Moran NA (2012) Establishment of char- under heat stress. Ecological Entomology, 27, 189–195.
acteristic gut bacteria during development of the honeybee Morales-Jimenez J, Z ~ iga G, Villa-Tanaca L, Hern
un andez-
worker. Applied and Environmental Microbiology, 78, 2830–2840. Rodrıguez C (2009) Bacterial community and nitrogen fixa-
Matsuura Y, Kikuchi Y, Hosokawa T et al. (2012) Evolution of tion in the red turpentine beetle, Dendroctonus valens LeConte
symbiotic organs and endosymbionts in lygaeid stinkbugs. (Coleoptera: Curculionidae: Scolytinae). Microbial Ecology, 58,
The ISME Journal, 6, 397–409. 879–891.
Mattson WJ Jr (1980) Herbivory in relation to plant nitrogen Morales-Jimenez J, Z ~ iga G, Ramırez-Saad HC, Hernandez-
un
content. Annual Review of Ecology and Systematics, 11, 119–161. Rodrıguez C (2012) Gut-associated bacteria throughout the
May RM (1988) How many species are there on Earth? Science, life cycle of the bark beetle Dendroctonus rhizophagus Thomas
241, 1441–1449. and Bright (Curculionidae: Scolytinae) and their cellulolytic
McClean APD, Oberholzer PCJ (1965) Citrus psylla, a vector of activities. Microbial Ecology, 64, 268–278.
the greening disease of Sweet Orange. South African Journal Morales-Jimenez J, Vera-Ponce de Le on A, Garcıa-Domınguez
of Agricultural Science, 8, 297–298. A et al. (2013) Nitrogen-fixing and uricolytic bacteria associ-
McCutcheon JP, Moran NA (2007) Parallel genomic evolution ated with the gut of Dendroctonus rhizophagus and Dendroct-
and metabolic interdependence in an ancient symbiosis. onus valens (Curculionidae: Scolytinae). Microbial Ecology, 66,
Proceedings of the National Academy of Sciences, USA, 104, 200–210.
19392–19397. Moran NA (2007) Symbiosis as an adaptive process and source
McCutcheon JP, Moran NA (2010) Functional convergence in of phenotypic complexity. Proceedings of the National Academy
reduced genomes of bacterial symbionts spanning 200 mil- of Sciences, USA, 104(Suppl 1), 8627–8633.
lion years of evolution. Genome Biology and Evolution, 2, Moran NA, Jarvik T (2010) Lateral transfer of genes from fungi
708–718. underlies carotenoid production in aphids. Science, 328,
McCutcheon JP, Moran NA (2012) Extreme genome reduction 624–627.
in symbiotic bacteria. Nature Reviews Microbiology, 10, 13–26. Moran NA, Plague GR, Sandstr€ om JP, Wilcox JL (2003) A
McCutcheon JP, von Dohlen CD (2011) An interdependent genomic perspective on nutrient provisioning by bacterial
metabolic patchwork in the nested symbiosis of mealybugs. symbionts of insects. Proceedings of the National Academy of
Current Biology, 21, 1366–1372. Sciences, USA, 100(Suppl 2), 14543–14548.

© 2013 John Wiley & Sons Ltd


1494 A . K . H A N S E N and N . A . M O R A N

Moran NA, Tran P, Gerardo NM (2005) Symbiosis and insect Pauchet Y, Wilkinson P, Chauhan R, Ffrench-Constant RH
diversification: an ancient symbiont of sap-feeding insects (2010) Diversity of beetle genes encoding novel plant cell
from the bacterial phylum Bacteroidetes. Applied and Environ- wall degrading enzymes. PLoS One, 5, e15635.
mental Microbiology, 71, 8802–8810. Peccoud J, Simon J-C, McLaughlin HJ, Moran NA (2009) Post-
Moran NA, McCutcheon JP, Nakabachi A (2008) Genomics and Pleistocene radiation of the pea aphid complex revealed by
evolution of heritable bacterial symbionts. Annual Review of rapidly evolving endosymbionts. Proceedings of the National
Genetics, 42, 165–190. Academy of Sciences, USA, 106, 16315–16320.
Moran NA, Hansen AK, Powell JE, Sabree ZL (2012) Perez-Brocal V, Gil R, Ramos S et al. (2006) A small microbial
Distinctive gut microbiota of honey bees assessed using genome: the end of a long symbiotic relationship? Science,
deep sampling from individual worker bees. PLoS One, 7, 314, 312–313.
e36393. Pilon FM, Vis^ otto LE, Guedes RNC, Oliveira MGA (2013) Pro-
Mueller UG (2012) Symbiont recruitment versus ant-symbiont teolytic activity of gut bacteria isolated from the velvet bean
co-evolution in the attine ant-microbe symbiosis. Current caterpillar Anticarsia gemmatalis. Journal of Comparative Physi-
Opinion in Microbiology, 15, 269–277. ology B, Biochemical, Systemic, and Environmental Physiology
Munson MA, Baumann P, Clark MA et al. (1991) Evidence for Feb 8. [Epub ahead of print]. doi:10.1007/s00360-013-0744-5.
the establishment of aphid-eubacterium endosymbiosis in an Poliakov A, Russell CW, Ponnala L et al. (2011) Large-scale
ancestor of four aphid families. Journal of Bacteriology, 173, label-free quantitative proteomics of the pea aphid-Buchnera
6321–6324. symbiosis. Molecular & Cellular Proteomics, 10, 110007039.
Nakabachi A, Yamashita A, Toh H et al. (2006) The 160-kilo- Prado SS, Golden M, Follett PA, Daugherty MP, Almeida RPP
base genome of the bacterial endosymbiont Carsonella. (2009) Demography of gut symbiotic and aposymbiotic
Science, 314, 267. Nezara viridula L. (Hemiptera: Pentatomidae). Environmental
Nichols-Orians C (1991a) Condensed tannins, attine ants, and Entomology, 38, 103–109.
the performance of a symbiotic fungus. Journal of Chemical Price DRG, Duncan RP, Shigenobu S, Wilson ACC (2011) Gen-
Ecology, 17, 1177–1195. ome expansion and differential expression of amino acid
Nichols-Orians C (1991b) Differential effects of condensed transporters at the aphid/Buchnera symbiotic interface.
and hydrolyzable tannin on polyphenol oxidase activity of Molecular Biology and Evolution, 28, 3113–3126.
attine symbiotic fungus. Journal of Chemical Ecology, 17, Raddadi N, Gonella E, Camerota C et al. (2011) “Candidatus
1811–1819. Liberibacter europaeus” sp. nov. that is associated with and
Nikoh N, Hosokawa T, Oshima K, Hattori M, Fukatsu T (2011) transmitted by the psyllid Cacopsylla pyri apparently behaves
Reductive evolution of bacterial genome in insect gut envi- as an endophyte rather than a pathogen. Environmental
ronment. Genome Biology and Evolution, 3, 702–714. Microbiology, 13, 414–426.
Nishikori K, Morioka K, Kubo T, Morioka M (2009) Age- and Raffa KF, Berryman AA (1983) The role of host plant resistance
morph-dependent activation of the lysosomal system and in the colonization behavior and ecology of bark beetles
Buchnera degradation in aphid endosymbiosis. Journal of (Coleoptera: Scolytidae). Ecological Monographs, 53, 27–49.
Insect Physiology, 55, 351–357. Redford AJ, Bowers RM, Knight R, Linhart Y, Fierer N (2010)
Nygaard S, Zhang G, Schiøtt M et al. (2011) The genome of the The ecology of the phyllosphere: geographic and phyloge-
leaf-cutting ant Acromyrmex echinatior suggests key adapta- netic variability in the distribution of bacteria on tree leaves.
tions to advanced social life and fungus farming. Genome Environmental Microbiology, 12, 2885–2893.
research, 21, 1339–1348. Reid NM, Addison SL, Macdonald LJ, Lloyd-Jones G (2011)
Oliver KM, Russell JA, Moran NA, Hunter MS (2003) Faculta- Biodiversity of active and inactive bacteria in the gut flora of
tive bacterial symbionts in aphids confer resistance to para- wood-feeding huhu beetle larvae (Prionoplus reticularis).
sitic wasps. Proceedings of the National Academy of Sciences, Applied and Environmental Microbiology, 77, 7000–7006.
USA, 100, 1803–1807. Reymond N, Calevro F, Viñuelas J et al. (2006) Different levels
Oliver KM, Moran NA, Hunter MS (2005) Variation in resis- of transcriptional regulation due to trophic constraints in the
tance to parasitism in aphids is due to symbionts not host reduced genome of Buchnera aphidicola APS. Applied and
genotype. Proceedings of the National Academy of Sciences, Environmental Microbiology, 72, 7760–7766.
USA, 102, 12795–12800. Richard F-J, Mora P, Errard C, Rouland C (2005) Digestive
Oliver KM, Degnan PH, Hunter MS, Moran NA (2009) Bacte- capacities of leaf-cutting ants and the contribution of their
riophages encode factors required for protection in a symbi- fungal cultivar to the degradation of plant material. Journal
otic mutualism. Science, 325, 992–994. of Comparative Physiology B, Biochemical, Systemic, and Environ-
Oliver KM, Degnan PH, Burke GR, Moran NA (2010) Faculta- mental Physiology, 175, 297–303.
tive symbionts in aphids and the horizontal transfer of Rio RVM, Wu Y-N, Filardo G, Aksoy S (2006) Dynamics of
ecologically important traits. Annual Review of Entomology, multiple symbiont density regulation during host develop-
55, 247–266. ment: tsetse fly and its microbial flora. Proceedings of the
Paganini J, Campan-Fournier A, Da Rocha M et al. (2012) Con- Royal Society B: Biological Sciences, 273, 805–814.
tribution of lateral gene transfers to the genome composition Robert-Seilaniantz A, Navarro L, Bari R, Jones JD (2007) Patho-
and parasitic ability of root-knot nematodes. PLoS One, 7, logical hormone imbalances. Current Opinion in Plant Biology,
e50875. 10, 372–379.
Paine TD, Raffa KF, Harrington TC (1997) Interactions among Robinson CJ, Schloss P, Ramos Y, Raffa K, Handelsman J
scolytid bark beetles, their associated fungi, and live host (2010) Robustness of the bacterial community in the cabbage
conifers. Annual Review of Entomology, 42, 179–206. white butterfly larval midgut. Microbial Ecology, 59, 199–211.

© 2013 John Wiley & Sons Ltd


E V A L U A T I N G S Y M B I O N T E F F E C T S O N H O S T P L A N T U S E 1495

Rohfritsch O (2008) Plants, gall midges, and fungi: a three-compo- Six DL, Wingfield MJ (2011) The role of phytopathogenicity in
nent system. Entomologia Experimentalis et Applicata, 128, 208–216. bark beetle–fungus symbioses: a challenge to the classic par-
Ryan CA (1990) Protease inhibitors in plants: Genes for adigm. Annual Review of Entomology, 56, 255–272.
improving defenses against insects and pathogens. Annual Slansky F, Scriber JM (1985) Food consumption and utilization.
Review of Phytopathology, 28, 425–449. In: Comprehensive Insect Physiology, Biochemistry and Pharma-
Sabree ZL, Huang CY, Okusu A, Moran NA, Normark BB cology (eds Kerkut GA, Gilbert LI), Vol. 4. pp. 87–163. Perg-
(2013) The nutrient-supplying capabilities of Uzinura, an amon Press, Oxford, UK.
endosymbiont of armoured scale insects. Environmental Sloan DB, Moran NA (2012a) Genome reduction and co-evolu-
Microbiology, 15, 1988–1999. tion between the primary and secondary bacterial symbionts
Sandstr€ om JP, Moran NA (1999) How nutritionally unbalanced of psyllids. Molecular Biology and Evolution, 29, 3781–3792.
is phloem sap for aphids? Entomologia Experimentalis et Appli- Sloan DB, Moran NA (2012b) Endosymbiotic bacteria as a
cata, 91, 203–210. source of carotenoids in whiteflies. Biology Letters, 8, 986–989.
Sandstr€ om JP, Moran NA (2001) Amino acid budgets in three Stoll S, Feldhaar H, Fraunholz MJ, Gross R (2010) Bacteriocyte
aphid species using the same host plant. Physiological Ento- dynamics during development of a holometabolous insect, the
mology, 26, 202–211. carpenter ant Camponotus floridanus. BMC Microbiology, 10, 308.
Sandstr€ om J, Pettersson J (1994) Amino acid composition of Stout MJ, Thaler JS, Thomma BPHJ (2006) Plant-mediated
phloem sap and the relation to intraspecific variation in pea interactions between pathogenic microorganisms and her-
aphid (Acyrthosiphon pisum) performance. Journal of Insect bivorous arthropods. Annual Review of Entomology, 51, 663–
Physiology, 40, 947–955. 689.
Scarborough CL, Ferrari J, Godfray HCJ (2005) Aphid pro- Strong DR, Lawton JH (1984) Insects on Plants: Community Patterns
tected from pathogen by endosymbiont. Science, 310, 1781. and Mechanisms. Harvard University Press, Cambridge, MA.
Schauer C, Thompson CL, Brune A (2012) The bacterial Suen G, Scott JJ, Aylward FO et al. (2010) An insect herbivore
community in the gut of the Cockroach Shelfordella lateralis microbiome with high plant biomass-degrading capacity.
reflects the close evolutionary relatedness of cockroaches and PLoS Genetics, 6, e1001129.
termites. Applied and Environmental Microbiology, 78, 2758–2767. Suen G, Teiling C, Li L et al. (2011) The genome sequence of
Schiøtt M, De Fine Licht HH, Lange L, Boomsma JJ (2008) the leaf-cutter ant Atta cephalotes reveals insights into its obli-
Towards a molecular understanding of symbiont function: gate symbiotic lifestyle. PLoS Genetics, 7, e1002007.
identification of a fungal gene for the degradation of xylan in Sugio A, Kingdom HN, MacLean AM, Grieve VM, Hogenhout
the fungus gardens of leaf-cutting ants. BMC Microbiology, 8, 40. SA (2011) Phytoplasma protein effector SAP11 enhances
Schiøtt M, Rogowska-Wrzesinska A, Roepstorff P, Boomsma JJ insect vector reproduction by manipulating plant develop-
(2010) Leaf-cutting ant fungi produce cell wall degrading ment and defense hormone biosynthesis. Proceedings of the
pectinase complexes reminiscent of phytopathogenic fungi. National Academy of Sciences, USA, 108, E1254–E1263.
BMC Biology, 8, 156. Sun BF, Xiao JH, He SM et al. (2013) Multiple ancient horizon-
Schoonhoven LM, van Loon JJA, Dicke M (2005) Insect-Plant tal gene transfers and duplications in lepidopteran species.
Biology. Oxford University Press, Oxford, UK. Insect Molecular Biology, 22, 72–87.
Schuster LN, Sommer RJ (2012) Expressional and functional Tamas I (2002) 50 Million years of genomic stasis in endosym-
variation of horizontally acquired cellulases in the nematode biotic bacteria. Science, 296, 2376–2379.
Pristionchus pacificus. Gene, 506, 274–282. Tan Q, Zhang L, Grant J, Cooper P, Tegeder M (2010)
Shen Z (2003) Polygalacturonase from Sitophilus oryzae: possible Increased phloem transport of S-methylmethionine positively
horizontal transfer of a pectinase gene from fungi to weevils. affects sulfur and nitrogen metabolism and seed develop-
Journal of Insect Science, 3, 24. ment in pea plants. Plant Physiology, 154, 1886–1896.
Shigenobu S, Watanabe H, Hattori M, Sakaki Y, Ishikawa H Tang X, Freitak D, Vogel H et al. (2012) Complexity and vari-
(2000) Genome sequence of the endocellular bacterial symbi- ability of gut commensal microbiota in polyphagous lepidop-
ont of aphids Buchnera sp. APS. Nature, 407, 81–86. teran larvae. PLoS One, 7, e36978.
Silva A, Bacci M, Pagnocca FC, Bueno OC, Hebling MJA Teixeira DDC, Saillard C, Eveillard S et al. (2005) “Candidatus
(2006a) Starch metabolism in Leucoagaricus gongylophorus, the Liberibacter americanus”, associated with citrus huanglongbing
symbiotic fungus of leaf-cutting ants. Microbiological Research, (greening disease) in S~ao Paulo State, Brazil. International Journal
161, 299–303. of Systematic and Evolutionary Microbiology, 55, 1857–1862.
Silva A, Bacci M, Pagnocca FC, Bueno OC, Hebling MJA Teixeira L, Ferreira A, Ashburner M (2008) The bacterial sym-
(2006b) Production of polysaccharidases in different carbon biont Wolbachia induces resistance to RNA viral infections in
sources by Leucoagaricus gongylophorus M€ oller (Singer), the Drosophila melanogaster. PLoS Biology, 6, e2.
symbiotic fungus of the leaf-cutting ant Atta sexdens Linna- Thao ML, Moran NA, Abbot P et al. (2000a) Cospeciation of
eus. Current Microbiology, 53, 68–71. psyllids and their primary prokaryotic endosymbionts.
Sittenfeld A, Uribe-Lorıo L, Mora M et al. (2002) Does a Applied and Environmental Microbiology, 66, 2898–2905.
polyphagous caterpillar have the same gut microbiota when Thao ML, Clark MA, Baumann L et al. (2000b) Secondary en-
feeding on different species of food plants? Revista de Biologıa dosymbionts of psyllids have been acquired multiple times.
Tropical, 50, 547–560. Current Microbiology, 41, 300–304.
Six DL (2012) Ecological and evolutionary determinants of bark Tsuchida T, Koga R, Fukatsu T (2004) Host plant specialization
beetle —fungus symbioses. Insects, 3, 339–366. governed by facultative symbiont. Science, 303, 1989.

© 2013 John Wiley & Sons Ltd


1496 A . K . H A N S E N and N . A . M O R A N

Urban JM, Cryan JR (2012) Two ancient bacterial endos- Watanabe H, Noda H, Tokuda G, Lo N (1998) A cellulase gene
ymbionts have coevolved with the planthoppers (Insecta: of termite origin. Nature, 394, 330–331.
Hemiptera: Fulgoroidea). BMC Evolutionary Biology, 12, 87. Weintraub PG, Beanland L (2006) Insect vectors of phytoplas-
Vega FE, Blackburn MB, Kurtzman CP, Dowd PF (2003) Identifica- mas. Annual Review of Entomology, 51, 91–111.
tion of a coffee berry borer-associated yeast: does it break down Weis AE (1982) Use of symbiotic fungus by the gall maker
caffeine? Entomologia Experimentalis et Applicata, 107, 19–24. Asteromyia carbonifera to inhibit attack by the parasitoid Tory-
nuelas J, Febvay G, Duport G et al. (2011) Multimodal
Vi~ mus capite. Ecology, 63, 1602–1605.
dynamic response of the Buchnera aphidicola pLeu plasmid to Wilson ACC, Ashton PD, Calevro F et al. (2010) Genomic
variations in leucine demand of its host, the pea aphid Acyr- insight into the amino acid relations of the pea aphid, Acyr-
thosiphon pisum. Molecular Microbiology, 81, 1271–1285. thosiphon pisum, with its symbiotic bacterium Buchnera
Vis^
otto LE, Oliveira MGA, Guedes RNC, Ribon AOB, Good- aphidicola. Insect Molecular Biology, 19(Suppl 2), 249–258.
God PIV (2009) Contribution of gut bacteria to digestion and Wolschin F, H€ olldobler B, Gross R, Zientz E (2004) Replication of
development of the velvetbean caterpillar, Anticarsia gemma- the endosymbiotic bacterium Blochmannia floridanus is corre-
talis. Journal of Insect Physiology, 55, 185–191. lated with the developmental and reproductive stages of its
Vogel KJ, Moran NA (2010) Sources of variation in dietary ant host. Applied and Environmental Microbiology, 70, 4096–4102.
requirements in an obligate nutritional symbiosis. Proceedings Wu D, Daugherty SC, Van Aken SE et al. (2006) Metabolic
of the Royal Society B: Biological Sciences, 278, 115–121. complementarity and genomics of the dual bacterial
Vorburger C, Gehrer L, Rodriguez P (2010) A strain of the bac- symbiosis of sharpshooters. PLoS Biology, 4, e188.
terial symbiont Regiella insecticola protects aphids against
parasitoids. Biology Letters, 6, 109–111.
Walters DR, McRoberts N, Fitt BDL (2008) Are green islands
red herrings? Significance of green islands in plant inter- N.A.M.’s research is on genome evolution and on
actions with pathogens and pests. Biological Reviews, 83, evolutionary and ecological associations between insects
79–102. and microorganisms. A.K.H. is interested in evolution-
Warnecke F, Luginb€ uhl P, Ivanova N et al. (2007) Metagenomic ary and ecological genomics and the molecular mecha-
and functional analysis of hindgut microbiota of a wood- nisms that underpin insect–microbe coevolution.
feeding higher termite. Nature, 450, 560–565.
Watanabe H, Tokuda G (2010) Cellulolytic systems in insects.
Annual Review of Entomology, 55, 609–632.

© 2013 John Wiley & Sons Ltd

You might also like