Truman 2019 The Evolution of Insect Metamorphosis

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Current Biology

Review

The Evolution of Insect Metamorphosis


James W. Truman
Department of Biology and Friday Harbor Laboratories, University of Washington, 620 University Road, Friday Harbor, WA 98250, USA
Correspondence: trumanj@janelia.hhmi.org
https://doi.org/10.1016/j.cub.2019.10.009

The evolution of insect metamorphosis is one of the most important sagas in animal history, transforming
small, obscure soil arthropods into a dominant terrestrial group that has profoundly shaped the evolution
of terrestrial life. The evolution of flight initiated the trajectory towards metamorphosis, favoring enhanced
differences between juvenile and adult stages. The initial step modified postembryonic development, result-
ing in the nymph–adult differences characteristic of hemimetabolous species. The second step was to com-
plete metamorphosis, holometaboly, and occurred by profoundly altering embryogenesis to produce a larval
stage, the nymph becoming the pupa to accommodate the deferred development needed to make the adult.
These changing life history patterns were intimately linked to two hormonal systems, the ecdysteroids and
the juvenile hormones (JH), which function in both embryonic and postembryonic domains and control the
stage-specifying genes Krüppel homolog 1 (Kr-h1), broad and E93. The ecdysteroids induce and direct molt-
ing through the ecdysone receptor (EcR), a nuclear hormone receptor with numerous targets including a
conserved transcription factor network, the ‘Ashburner cascade’, which translates features of the ecdyste-
roid peak into the different phases of the molt. With the evolution of metamorphosis, ecdysteroids acquired
a metamorphic function that exploited the repressor capacity of the unliganded EcR, making it a hormone-
controlled gateway for the tissue development preceding metamorphosis. JH directs ecdysteroid action,
controlling Kr-h1 expression which in turn regulates the other stage-specifying genes. JH appears in basal
insect groups as their embryos shift from growth and patterning to differentiation. As a major portion of
embryogenesis was deferred to postembryonic life with the evolution of holometaboly, JH also acquired a
potent role in regulating postembryonic growth and development. Details of its involvement in broad expres-
sion and E93 suppression have been modified as life cycles became more complex and likely underlie some
of the changes seen in the shift from incomplete to complete metamorphosis.

Introduction for hormone extraction and chemical characterization. Ensuing


Amphibians and insects provide most people with their first genetic and molecular studies, though, gradually shifted to
contact with metamorphosis. The striking transformation of a Drosophila melanogaster, which still serves as an unrivaled plat-
tadpole into a frog is relatively rapid but one can track its form for gene discovery. However, many derived features of
progression from day to day as its limbs extend and its tail is re- Drosophila, such as an invariant number of larval instars and
sorbed. Metamorphosis in insects, though, seems more myste- an almost complete replacement of larval cells by those of the
rious as their external exoskeleton hides the gradual changes adult at metamorphosis, have resulted in a de-emphasis or
occurring inside, and one only sees the results when the old loss of some aspects of the hormonal control seen in less derived
outer cuticle is abruptly shed, such as when the butterfly insects. Consequently, despite the fact that work on Drosophila
emerges from its chrysalis. Metamorphosis is an ancient life his- is responsible for the discovery of virtually all the genes involved
tory trait that extends deep to the roots of the amphibians and in hormone reception, their downstream effects, and stage spec-
many groups of invertebrates. Insects are unusual, though, in ificity of action, the broader functions of such genes often re-
that their metamorphosis is a derived trait that evolved well after mained obscure for decades and were finally shown in insects
they colonized the land. Insect orders approximating different with less derived life history patterns. For example, work on
steps in the evolution of metamorphosis exist today, so that the flour beetle, Tribolium castaneum, provided the proof that
comparative studies allow a piecing together of the molecular, the Methoprene tolerant gene (Met) does, indeed, encode the re-
developmental and endocrine changes that accompanied this ceptor for juvenile hormone (JH), a key hormone that suppresses
transition. metamorphosis [1,2]. Also, studies on this beetle and the cock-
Insect metamorphosis is under hormonal control, which has roach Blattella germanica showed that the helix-turn-helix tran-
been a subject of experimental study for almost a century. The scription factor Eip93F (E93) is a master stage-specifying gene
importance of insects as competitors for food and fiber as well that stands on top of a gene cascade that produces the adult
as their impact on human and animal health spurred extensive phenotype [3].
research into their physiology and development, and especially Besides Drosophila, this discussion draws on other insect
into how they control their metamorphosis. The early stages of model systems such as Bombyx mori, Manduca sexta, Tribolium
this research typically used large insects which facilitated surgi- castaneum, Oncopeltus fasciatus, Pyrrhocoris apterus and
cal manipulations and the isolation of sufficient amounts of tissue Blattella germanica, for which genomic information and gene

R1252 Current Biology 29, R1252–R1268, December 2, 2019 ª 2019 Elsevier Ltd.
Current Biology

Review

Ametabolous Hemimetabolous Holometabolous

Palaeoptera
Palaeopte
era Polyneoptera Paraneoptera

Mantophasmatodea
smatodea
Ephemeroptera
ptera
natha
Archaeognatha

Grylloblattidae
idae

Raphidioptera
tera
Phasmatodeadea

tera
Thysanoptera

Hymenoptera
era

era
Siphonaptera
era
Megaloptera
Phthiraptera
ra
Dermaptera ra

Lepidoptera
ra
Embioptera a

Coleoptera
a

a
Trichoptera
a
Zygentoma

a
Neuroptera
Plecopteraa
Orthopteraa

a
Mecoptera
Hemiptera
a
Psocodea
Mantodea
Zoraptera

Blattodea
Odonata

Diptera
0
N

66
Age [millions of years ago]

145

200

250
300

360
Holometabola
420
Pterygota
485
Current Biology

Figure 1. Phylogeny of insects showing the major types of development.


The figure shows how the three major types of insect development, ametabolous, hemimetabolous and holometabolous, map onto insect phylogeny, with
examples of the immature and adult stages for each. The asterisk indicates neometabolous forms that have independently evolved a life history with a larva, pupa
and adult. For the major orders the width of the boxes show the number of families through time. Phylogeny from [134,135] with family distributions adapted from
[14] based on [136].

suppression techniques can be woven into a rich history of endo- process termed ecdysis. Consequently, the growth of the juve-
crine experimentation. Although these few systems do not begin nile alternates between a period of feeding and growth (the inter-
to cover the amazing range of variation that one sees in the in- molt) and a molt. The span from one ecdysis to the next is termed
sects, it is heartening that highly derived life history strategies, an instar. In most insects, there is a characteristic number of in-
such as female neoteny in Strepsiptera [4,5] and scale insects stars between hatching and metamorphosis, but the decision to
[6] or the complete metamorphosis that evolved in parallel in begin metamorphosis typically depends on reaching a charac-
thrips [7], utilize the same molecular switches that regulate teristic species-specific size [8]. Therefore, suboptimal nutrition
more traditional life history patterns. This review explores the or injuries or disease may result in the intercalation of additional
developmental changes that accompanied the evolution of the molts until this size is achieved. Insects are extremely diverse in
insect larval and pupal stages. It then focuses on the two main their feeding strategies, however, and in some species the larva
hormonal systems that regulate insect molting and metamor- can opt for an earlier instar to start metamorphosis if food condi-
phosis and considers how these systems evolved to support tions are poor (for example [9]).
the increasing complexity of insect life histories. Finally, it exam- By any criteria, insects are a spectacularly successful terres-
ines our growing knowledge of the function of the stage specifi- trial group with diverse life history strategies and 28 extant orders
cation genes that regulate the phenotype of each major life stage (Figure 1). The true insects and other hexapods such as spring-
and how these genes and their control have changed with the tails (Collembola) and diplurans (Diplura) diverged from their
increasing complexity of insect life histories. aquatic crustacean ancestors and were part of the arthropod in-
vasion of land that began over 450 million years ago. The life his-
Insect Life History Strategies tory of these earliest insects was similar to those of the jumping
Because their rigid exoskeleton constrains growth and changes bristletails (Archaeognatha) and silverfish (Zygentoma) of today.
in morphology, an insect’s life history is punctuated by molts, These primitively wingless insects have an ametabolous lifestyle
during which a new cuticle is formed and the old one shed, a which shows direct development: the hatchling is a miniature

Current Biology 29, R1252–R1268, December 2, 2019 R1253


Current Biology

Review

version of the adult and its morphology is essentially unchanged affecting adult specializations for dispersal and reproduction
during its growth to an adult. Like most other arthropods, the [14]. This life history innovation resulted in the rapid diversifica-
adults in these ametabolous orders continue to molt, typically tion and radiation within the Holometabola [15], which includes
alternating a molt with a bout of reproduction [10]. the orders that have the most species (Coleoptera), the most
The insects are the only invertebrates to have evolved pow- biomass (Hymenoptera), and the greatest impacts on agriculture
ered flight and they owned the skies for over 100 million years un- (Lepidoptera) and on human and animal health (Diptera).
til the Pterosaurs appeared in the late Triassic. Flight provided a In holometabolous insects, the transition from larva to the
means to avoid predation, to move rapidly between dispersed adult takes place through the nonfeeding, transitional stage
food patches, and to undertake long-range dispersals. Early or- known as the pupa. The oldest unequivocal fossils of holometab-
ders of winged insects, such as the extinct Palaeodictyoptera olous larvae are from about 320 million years ago [16]. Curiously,
and Megasecoptera, were also considered ametabolous in this era of giant insects, the earliest holometabolans were
because their juveniles had small articulated winglets that even- quite small. Their small size may have allowed them to exploit
tually became large enough to support flight in later instars niches that were unavailable or not productive for their larger
[11,12]. The evolution of wings and flight, though, posed devel- relatives. A drop in ancient oxygen levels may have subsequently
opmental problems, the solution for which provided the opportu- doomed the giants but would have had little effect on the
nity for the first step into metamorphosis [13]. Instead of growing small holometabolous forms and may have opened the door
as small articulated winglets, wing growth in juveniles took place to their radiation [16]. Some hemimetabolous groups, such
in immobile wing pads, which then morphed into articulated as thrips and some hemipterans, have independently evolved a
wings at the adult molt. A second, wing-related change evident holometabolous-like lifestyle, a condition sometimes called
in today’s winged insects (the Pteryogota) was that the adult ‘neometaboly’ [17].
molt became a terminal molt. The cessation of adult molting
was likely related to making the wing as light as possible. Weight The Larva of Insects with Complete Metamorphosis
reduction occurs by the death of the epidermal cells that secrete Ideas about insect metamorphosis extend back to Aristotle
the cuticle of the wing membrane after adult emergence. How- [18,19] and were later built on by the anatomist Sir William Har-
ever, without these cells, the cuticle of the wing membrane vey, who considered the larva as the product of an ‘imperfect
cannot be made again. The only extant insects that molt their egg’, from which the animal hatched before acquiring its spe-
wings are the mayflies (Ephemeroptera), a very basal order cies-typical form. The pupa then served as a second egg in
which has a winged subimago stage that quickly molts to the which the proper form of the animal was constructed from the re-
winged adult. The subimagos have thick wings and are very mains of the larva. Later refinement of these ideas held that the
awkward flyers. Interestingly, the loss of adult molting seems pupa corresponded to the nymph of hemimetabolous forms
to be irreversible: two parasitic orders of insects, the lice (Pthir- [20,21], but these ideas for the correspondence of the life stages
aptera) and the fleas (Siphonaptera), are wingless as adults but in holometabolous and hemimetabolous insects have been
neither have reacquired adult molting. challenged by other researchers who considered larvae and
The evolution of flight, then, ushered in the strategy of incom- nymphs to be equivalent stages [22,23]. They held that the
plete metamorphosis, also known as hemimetabolous develop- pupa is a modified last nymphal stage [22] or that the last
ment. The immature stage is called a nymph, and nymphs nymphal molt splits into a two-step process with the pupa form-
typically resemble the adult but lack wings. External wing pads ing the first step [23]. Discussions of these latter ideas can be
appear during nymphal growth, but they do not transform into found in other papers [24,25], but we found that the idea of the
functional wings until the adult molt. Nymphs and adults usually larva arising from a modified embryonic stage fits better with
occupy similar habitats, exploit similar food sources, but often the available developmental, endocrine and molecular data
have different cuticle types. While nymphs and adults usually [26–28]. I will use the terms ‘larvae’ and ‘nymphs’ for the imma-
have similar morphologies, in orders like the Odonata (dragon- ture feeding stages of holometabolous and hemimetabolous in-
flies) and Ephemeroptera (mayflies) that have aquatic nymphs sects, respectively.
that transform into aerial adults, the two stages may show During embryogenesis, regions of the embryo organize into
marked differences in body form. The high mobility of the hemi- primordia destined to form legs, antennae, eyes and so on. In
metabolous adult likely set the stage for the step to complete hemimetabolous insects, such primordia grow and pattern
metamorphosis. themselves as they form miniature versions of the adult struc-
Evolution of the hemimetabolous lifestyle primarily involved ture. The embryos of holometabolous species have the same
alterations in postembryonic development. As in their set of primordia, but clonal experiments with Drosophila demon-
ametabolous ancestors, the basic adult body plan was laid strated that only part of each primordium is utilized to make a
down during embryogenesis, although some degree of delayed reduced, larval structure, while the remainder is set aside to
development, such as adding new rows of visual units (omma- contribute later to the adult counterpart [29]. This pattern is nicely
tidia) to the growing eye, occurs in some groups. The shift to illustrated for the compound eye (Figure 2A) [30,31]. In hemime-
holometabolous development, though, involved a fundamental tabolous embryos, the eye primordium begins organization with
alteration of embryonic development to produce a larval stage a row of ommatidia on its posterior border, followed by addition
that bore little or no resemblance to the adult. Adaptations of of successive rows as a morphogenetic wave progresses
the larval and adult stages were developmentally uncoupled, anteriorly across the primordium. The nymph hatches with a
which allowed the evolution of a larval stage with reduced well-developed compound eye, although some species retain
motility but enhanced capacity to exploit food resources without a primordium along the anterior margin to add additional rows

R1254 Current Biology 29, R1252–R1268, December 2, 2019


Current Biology

Review

A Embryo
Nymph Adult
Grasshopper

Hatch
Birth photoreceptors

0 50 100%
E1 Growth
E2
E3
Embryo Larva Pupa Adult
Moth

50 Growth* Morphogenesis Differentiation


0 E1 100%
E2 pp
E3
Day 0 Day 4 Isomorphic Morphogenetic
B C Growth: phase phase
p *
Eye ** Molt

Primordium size
* *
* *
0h
0h Fee
e
eed
ed 4d
ed
Feed 4d
Wing
Eye
Wing
L3 L4 L5 Pupa

Current Biology

Figure 2. Modification of embryonic and postembryonic development associated with the evolution of the larva.
(A) Comparison of eye development in a generalized hemimetabolous insect (grasshopper) with that in a holometabolous insect (moth). The embryonic timeline
(% of development) shows the time of deposition of the three embryonic cuticles (black line: epicuticle; red line: procuticle). Photoreceptors start being born at
similar times, but moth embryos make only the first few photoreceptors while grasshoppers show extended photoreceptor production. The unused portion of the
eye primordium (light orange) in the caterpillar is maintained through larval growth and eventually forms the eye imaginal disc that forms the adult compound eye.
Adapted from [28] and based on [31,47]. Timing of embryonic molting [38,52]. (B) Caterpillars of Manduca sexta have some imaginal primordia that invaginate
early to form an imaginal disc (the wing disc, in green) and do not contribute to the larva, while others (red) are associated with their larval counterpart and make
larval cuticle but in the last larval instar they dedifferentiate and invaginate to make a late-forming imaginal disc. The confocal images of propidium iodide stained
whole mounts showing the eye primordium (p, dotted outline) at day 0 of the last larval instar when it is part of the epidermal monolayer and at Day 4 when it has
invaginated to form the eye disc; the asterisk indicates larval photoreceptors just behind the primordium. Through the same period the wing disc transforms
from a simple sack to a complexly folded structure with incipient wing veins. Adapted from [41]. (C) Imaginal primordia that become early-forming (green) versus
late-forming (red) discs show different patterns of isomorphic growth: as an early forming disc they grow through both the intermolt and molt periods while in the
late-forming strategy growth occurs only in the early phase of each larval molt. Under both strategies growth shifts into a morphogenetic mode early in the final
larval instar.

at subsequent nymphal molts. In embryos of holometabolous of the embryonic leg primordium [36,37], which undergoes
species, by contrast, the eye primordium makes only an initial, extensive proliferation and morphogenesis in the last larval stage
posterior set of photoreceptors which differentiate into the sim- and combines with some of the larval cells as they complete the
ple lateral eyes (stemmata) characteristic of larvae. The anterior patterning program to make the adult walking leg [37].
portion of the primordium remains dormant during larval growth The larval form, then, arises from the arrest of ancestral pat-
but, in the last larval instar, undergoes morphogenesis into the terns of tissue growth and patterning. The structures that are
eye imaginal disc [31]. organized prior to the arrest are adapted to the larval form, while
The legs of Lepidoptera present a more complex metamorphic the remainder of the primordium is carried forward to contribute
challenge, because the animal makes two different types of to the adult system (Figure 2A). The point at which the embryonic
complex leg during its lifetime: the caterpillar has shortened programs arrest to form the larval structure and the degree to
thoracic legs modified for grasping, while the adult moth or but- which larval cells then contribute to the adult version varies ac-
terfly has typical walking legs. The formation of a walking leg, cording to the species and organ system considered [27].
either during embryogenesis in a cricket [32,33] or postem- Insect embryos typically make three embryonic cuticles, and
bryonically in a Drosophila leg imaginal disc, occurs through a these are well developed in hemimetabolous forms [38]. The first
similar pattern of recruitment of proximodistal patterning genes embryonic cuticle is delicate and is deposited with the comple-
[34]. This patterning cascade is arrested at an intermediate point tion of segmentation and establishment of the embryonic
in embryos of the moth Manduca sexta, producing the modified primordia. The embryo then enters a phase of rapid growth
caterpillar leg [35]. The caterpillar leg, though, retains a remnant and patterning which is largely finished at dorsal closure when

Current Biology 29, R1252–R1268, December 2, 2019 R1255


Current Biology

Review

it deposits a second embryonic cuticle, called the pronymphal A more derived condition is seen when the imaginal primor-
cuticle. This pronymph stage has the general shape of the dium does not contribute to the larva but invaginates into the
nymph, although its proportions may be constrained by the con- body, where it becomes an early forming imaginal disc. Early
fines of the eggshell. Its cuticle differs from nymphal cuticle and forming discs are not found in the beetle Tribolium, although
may bear specializations for escape from the shell and the ovipo- they are found in more derived groups of beetles [26]. In Lepi-
sition site. Then follows tissue differentiation and maturation and doptera, the only early forming discs are the wing and genital
the deposition of the first nymphal cuticle. We proposed that the discs (Figure 2B), while in higher Diptera, such as Drosophila,
pronymph was the forerunner of the larval stage [26] because the virtually all adult structures come from early forming discs [29].
patterning processes that were interrupted in making the larva The only epidermal cells retaining an ancestral pattern of devel-
were those that occur during pronymph formation. In addition, opment are the abdominal histoblasts, small collections of
some pronymphs show unique behavioral adaptations, such diploid cells that make larval and then pupal cuticle before they
as body crawling movements that allow them to escape from proliferate and spread over the abdomen to form the adult
the oviposition site [39], and such stage-specific specializations cuticle.
may have been important pre-adaptations for the evolution of the The imaginal primordia show two different phases of growth
larva [26]. We further argued that the embryos of holometabolans (Figure 2B,C). In the preterminal larval instars, their growth is
underwent only two molts during embryogenesis [26]; however, sensitive to nutritional conditions and proportional to that of
subsequent work [38] showed that a number of holometabolous the larva as a whole. Through this period the early-forming discs
insect species also make three embryonic cuticles, although the have a distinct growth advantage over their late-forming coun-
second cuticle is somewhat reduced and lacks endocuticle. terparts (Figure 2C). At best, they secrete only a thin epicuticle
[40] and can proliferate through both the molt and intermolt pe-
The Pupa and Imaginal Discs and Primordia riods. The primordia that transform into discs in the last larval
As described above, the origin of the pupal stage has been a instar, however, are integral parts of the larval epidermis in pre-
controversial issue over the years. In support of the classic terminal instars. As such, they make cuticle through the latter
view that the pupa corresponds to the hemimetabolous nymph, part of the molt and the intermolt and their proliferation is
we proposed [26] that the immature stage responsible for growth restricted to a brief burst at the start of each molt. In the last
shifted from the nymph to the pronymph as the latter trans- instar both types of primordia shift to a morphogenetic program
formed into a larva that was capable of feeding. The number of that renders their growth independent of external nutrient input
pronymphal/larval instars increased to accommodate growth, [41,42].
while the number of nymphal instars decreased as the nymph
became a nonfeeding transitional stage, now recognized as Ecdysteroids and the Control of Molting and
the pupa. The pupa then provided a mold or template for the Metamorphosis
body of the adult, in which the adult structures subsequently The ecdysteroids are polyhydroxylated steroids that are made
differentiated and matured, but it also developed adaptations from dietary cholesterol or related plant sterols [43] (Figure 3A).
to ensure its survival, such as a characteristic cuticle type. If pu- Their primary function is to induce molting and they serve this
pation occurs within a protected environment, then the pupal function throughout the arthropods [44,45]. In insects, circu-
cuticle is typically thin, transparent and unadorned; pupae lating ecdysteroids typically come from the prothoracic glands,
exposed to the environment, such as the chrysalis of butterflies, although in some species other tissues such as the epidermis,
have a hardened cuticle that may be adorned with extensions the oenocytes, and the adult gonads also have this capacity.
and coloration that aid in its camouflage or protect it from attacks The prothoracic glands secrete ecdysone which is then further
by other arthropods. hydroxylated to 20-hydroxyecdysone (20E) by ecdysone 20-
The pupa is typically formed from reprogrammed larval cells monooxygenase in peripheral tissues [43,46]. 20E is normally
as well as cells from imaginal primordia (for example [37]). The considered to be the active form of the hormone, but ecdysone
relative contribution from these two depends on the organ and can function in its own right early in molting or metamorphic
species concerned. An imaginal primordium is a collection of programs (for example [47]). The control of ecdysone secretion
larval cells with a latent embryonic capacity that allows them to from the prothoracic glands is complex, but the two major
partially dedifferentiate and undergo growth and morphogenesis players are peptide hormones from brain neurosecretory cells:
to form part or all of an adult structure. Imaginal primordia are the prothoracicotropic hormone and the insulin-like peptides
usually associated with the corresponding larval structures, but [48]. The functions of ecdysteroids in the control of insect
their cells remain diploid while the surrounding larval cells may metamorphosis have striking parallels with those of the thyroid
become highly polyploid. The ancestral condition for such hormones in directing the metamorphosis of amphibians
primordia is likely similar to that seen for the antennal, leg and [49,50]. These parallels will be emphasized throughout this
eye primordia of Lepidoptera (Figure 2B,C) [27,40]. These section.
primordia are integral parts of the larval epidermis, making a As seen in Figure 3A, ecdysteroid levels are low through the in-
new larval cuticle at each molt and adding layers of endocuticle termolt but show a major peak during each molt. As ecdysone
through the intermolt. Early in the last larval instar, though, they 20-monooxygenase is upregulated during molts [43,46], periph-
stop making endocuticle, detach from the larval cuticle, partially eral tissues rather than the prothoracic glands determine the
dedifferentiate and initiate proliferation and morphogenesis [41]. steroid composition during the peak. Typically, ecdysone pre-
As they grow, they often invaginate into the body cavity to dominates at the start of the peak with 20E becoming predomi-
become late-forming imaginal discs. nant later (for example [51]; Figure 3A).

R1256 Current Biology 29, R1252–R1268, December 2, 2019


Current Biology

Review

A PTTH Peripheral tissues B


insulins
DNA Ligand
Brain Ecdysone 20 EcR-A A/B binding binding
Prothoracic
Glands monooxygenase Activation C D E F
EcR-B A/B Activation
20E
E
Ecdysteroid titer

9-cis RA
Other EcR RXR or EcR EcR

Coleoptera
EcR RXR or EcR EcR

Lepidoptera lipid
L4 L5 Pupa Adult
Days & diptera EcR USP
Co-repressors Co-activators
C 20E 20E
Larval tissues EcR USP EcR USP
EcR-B1 B1 B1
EcRE EcRE
Closed chromatin Open chromatin

Co-repressors Co-activators Other signaling pathways?


20E 20E
EcR USP EcR USP
EcR-A A A
EcRE EcRE RE
Imaginal tissues
Closed chromatin Open chromatin
Current Biology

Figure 3. The ecdysteroid system of insects.


(A) Ecdysone (E) is made by the prothoracic glands and secreted mainly under the influence of two brain hormones, the prothoracicotropic hormone (PTTH) and
the insect insulins. E is then converted to 20-hydroxyecdysone by peripheral tissues. The ecdysteroid titer is based on Manduca sexta [51,137] and shows total
ecdysteroids for the larval and pupal molts but ecdysone and 20E separately for the molt to the adult. (B) The structure of the ecdysone receptor (EcR) showing the
different domains of the protein. Most species express two EcR isoforms that vary in the transactivation ability of their A/B domain [58]. Depending on insect
group, EcR can either act as a homodimer or heterodimerize with the retinoid X receptor (RXR). RXR binds ligands such as 9-cis-retinoic acid (9-cis-RA) but there
are variants that lack the ligand binding domain or bind a constitutive structural lipid. The latter form is called Ultraspiracle (USP). The EcR in the latter group
cannot form homodimers. Adapted from [58]. (C) Prior to metamorphosis in Drosophila, larval and imaginal tissues express different EcR isoforms. In both cases,
the unliganded EcR–USP complex binds to its response element (EcRE) and binds to corepressors that suppress transcription. Ecdysteroid binding leads to
opening of chromatin structure and gene activation. The two major isoforms differ in this capacity and this difference may support different functions for ec-
dysteroids in larval versus imaginal tissues. EcRE, ecdysone response element; RE, response element for other transcription factors.

The deposition of a new cuticle, be it during the embryonic or As with other members of this nuclear receptor subfamily,
postembryonic period, requires a pulse of ecdysteroids. Each of some insect EcRs can homodimerize or form heterodimers
the three embryonic cuticles produced during grasshopper with another nuclear receptor, the retinoid X receptor (RXR). In
embryogenesis is associated with an ecdysteroid peak [52]. In more basal insects RXR binds fatty acids such as 9-cis-retinoic
Drosophila, embryos that lack the ecdysone biosynthetic en- acid, as seen in its vertebrate homolog. EcR–RXR has under-
zymes go through embryogenesis but do not make a larval gone considerable evolution within the insects (Figure 3B) [58].
cuticle and have a ghostly appearance that resulted in such The ligand-binding domain of RXR was apparently lost in bee-
mutants being called the ‘Halloween Group’. This collection of tles. In the Diptera and Lepidoptera, RXR has been modified
mutants subsequently provided the tools for dissection of the to bind a structural lipid and is called Ultraspiracle (USP), while
ecdysone biosynthetic pathway [53] EcR has evolved an expanded dimerization interface that
The Ecdysone Receptor facilitates making EcR–USP dimers but prevents formation of
Both ecdysteroids and thyroid hormones act through members EcR homodimers [58]. Because most of the model systems for
of the Group II subfamily of nuclear hormone receptors: the thy- studying ecdysteroid action come from these two orders, we
roid hormone receptor (TR) in the case of amphibians, and the know a lot about the functions of EcR–USP heterodimers, but
ecdysone receptor (EcR) in insects [54]. The structure of EcR is little about the EcR homodimers or effects of ligand binding
typical for this subfamily (Figure 3B): it has two C4 zinc fingers to RXR.
in the carboxyl domain to mediate DNA binding, the ecdyste- The EcR–USP heterodimer is typically in the nucleus and
roid-binding ‘E’ domain, and an amino-terminal A/B region that bound to DNA (Figure 3C) [59]. Without ligand, EcR–USP recruits
is differentially spliced to make two major isoforms, EcR-A and nucleosome remodeling proteins to compact the chromatin and
EcR-B [55,56]. These two isoforms are found throughout the in- cause local repression. With binding of 20E, coactivators such as
sects [57], but it is not known if their functions differ with the histone acetylases are exchanged for the co-repressors to
different types of metamorphic development. ‘open’ chromatin structure and facilitate transcription. The

Current Biology 29, R1252–R1268, December 2, 2019 R1257


Current Biology

Review

A Intermolt Molt component at the start of the ecdysteroid pulse [51], it is likely the
form responsible for these events.
Ecdysteroid titer

Preparatory High Ecdysial This is followed by the cuticle induction phase, starting with
phase 20E phase secretion of an outer epicuticle and then the procuticle made
Decline of protein and chitin. Whether the cuticle will be rigid or flexible
E or Cuticle
low 20E of 20E depends on its protein composition. The outer layers of the pro-
induction cuticle (the exocuticle) eventually become heavily crosslinked,
phase
making them extremely hard and resistant to degradation. The
inner layers (the endocuticle) continue to be deposited even after
the old cuticle is shed and the new one expanded. Ecdysone or
low levels of 20E cannot evoke cuticle production; rather, high
B 20E levels of 20E are required (for example [47]). This difference is
High 20E important, as the peripheral tissues are responsible for the con-
version of ecdysone to 20E [43], thereby controlling the length of
Low 20E Early genes
(High threshold) the preparatory phase and the onset of cuticle production. An
(or E?)
E74A, E75B extended preparatory phase is especially important for the
Early/late
molt from the pupa to the adult to allow sufficient time for the dif-
DHR3, DHR4
Early genes ferentiation and growth needed to make many adult structures.
(Low threshold)
Late The premature termination of this phase and precocious cuticle
e.g. E74B
βFtz-F1 production are seen when lepidopteran pupae are injected with a
Current Biology
high dose of 20E or other ecdysteroid agonists [65]: they show a
‘hyperecdysonism’ response, becoming malformed adults with
Figure 4. The relationship of phases of the molting process with the
naked cuticle and small eyes because the high 20E forced cuticle
ecdysteroid titer and the various components of the ‘Ashburner deposition before the completion of scale formation and eye
cascade’. growth.
(A) The ecdysteroid requirements for the three major phases of the molting
The ecdysial phase is the last one of the molt and involves the
process. (B) The components of the transcription factor cascade that are
active during the corresponding phases of the molt and their relationship to the degradation and shedding of the old cuticle and expansion and
ecdysteroid titer. The genes are examples of transcription factors from each tanning of the new one. During this phase, proteases and
class. chitinases degrade the old endocuticle, leaving the highly
cross-linked, but thin exocuticle. A cascade of neuropeptides
ligand-binding domain mediates core repression or activation then orchestrates a behavioral sequence to rupture and shed
based on whether or not ligand is present. The two major EcR the old cuticle (ecdysis sensu stricto), and then expand and
isoforms differ in both their functionality and their tissue distribu- crosslink the new one [66]. These events require the withdrawal
tion. In Drosophila melanogaster, EcR-B1 has a ligand-indepen- of ecdysteroids. This requirement for the withdrawal of 20E is
dent activation domain that is lacking in EcR-A. EcR-B1 is found why insects treated with ecdysteroid mimics that are not readily
through the larval stages with EcR-A joining it in the last instar metabolized typically die late in the molt without shedding their
and into metamorphosis [56,60], but at the latter time they are old cuticle [67].
separated into different cell types with larval cells expressing The different phases of the molt have characteristic patterns of
EcR-B1 and imaginal cells only EcR-A [56]. gene expression (Figure 4B). These were first identified in the
Ecdysteroids and Growth and Molting context of ecdysone’s effects on metamorphosis using cultured
The action of ecdysteroids in causing molting of the cuticle is an salivary glands of the midge Chironomous tentans [68]. Addition
ancient function of these hormones that precedes their involve- of ecdysone to larval glands evoked a stereotyped temporal
ment in metamorphosis. The insect cuticle is multilayered and sequence of ‘puffs’ — regions of expanded chromatin denoting
made primarily of protein and fibers of chitin, a polymer of transcriptional activity — along their giant polytene chromo-
N-acetyl-glucosamine [61,62]. Even in the largest insect, the somes, providing the first evidence that steroid hormones might
epidermis that secretes the cuticle is only a single cell thick. directly activate gene transcription. Comparison of puffing pat-
This monolayer produces the cuticle that lies above it. terns with and without inhibitors of protein synthesis indicated
The process of molting the cuticle can be divided into a that some genes were direct steroid targets while others were
number of phases [63] which I have condensed to three. downstream of this initial set. Quantitative analysis of these puffs
These phases are orchestrated by different features of the ec- and their steroid requirements by Michael Ashburner and col-
dysteroid pulse (Figure 4A). The rising ecdysteroid titer evokes leagues [69], using Drosophila salivary glands, identified a
the preparatory phase that involves a commitment component network of gene interactions often called the ‘Ashburner
which determines the nature of the molt (for example larval cascade’. While many of these genes are tissue specific and
versus pupal [64]) and the cellular events of detachment of the encode effector genes, there is a core set of transcription factors
epidermis from the overlying cuticle (apolysis), the secretion of that are expressed in all tissues, at each molt, and across spe-
molting fluid into the intervening space, cell division, and the dif- cies [70–72]. Many of these transcription factors are orphan nu-
ferentiation of new epidermal organelles such as a sensory hair clear receptors and they constitute a genetic circuit that interacts
and its socket. These events can be evoked by either low levels with and interprets that ecdysteroid titer, translating it into tem-
of 20E or ecdysone (for example, [47]). As ecdysone is the major poral programs of gene activity needed for the different phases

R1258 Current Biology 29, R1252–R1268, December 2, 2019


Current Biology

Review

of the molt [72]. Not all of the details of the interactions in this cir- switch to morphogenetic growth and pupal patterning [42].
cuit have been resolved, but enough is known about interactions This commitment to metamorphosis is irreversible and it may
of individual components to appreciate its overall function. occur in a mosaic fashion at different times across the tissues
As seen in Figure 4B, different parts of the cascade are asso- of the larva.
ciated with the different phases of the molt. The early genes, In Drosophila, this period also involves a dramatic increase in
which are directly induced by ecdysteroids, include low- EcR, but by the start of metamorphosis, the imaginal discs and
threshold and high-threshold classes [73]. The low-threshold the larval cells express different receptor isoforms [56]. These
class genes, such as E74B, appear to be associated with the different receptor isoforms reflect different roles for ecdysteroids
preparatory phase whereas the high threshold class genes, like in the metamorphosis of these two classes of cells (Figure 3C).
E74A and E75B, are associated with the cuticle initiation phase. Larval tissues show strong gene activation to initiate metamor-
Members of the high-threshold class also suppress the low- phosis, and the lack of ecdysteroids or EcR or the presence of
threshold genes, thereby terminating the preparatory phase. a dominant-negative form of EcR (EcRDN) [83] prevents
Consequently, the length of the preparatory phase is relatively metamorphosis. These tissues express EcR-B1, which has an
plastic and can be maintained by the appropriate hormone con- additional activation domain that supports high levels of gene in-
ditions as long as the high threshold genes are not induced. The duction in response to 20E [59]. The metamorphic development
high steroid titers also evoke the activation of early-late genes, of imaginal discs, by contrast, is blocked by the lack of ecdyste-
such as Drosophila Hormone Receptor 3 (DHR3), that have roids or the presence of EcRDN, but occurs if EcR is removed,
some steroid sensitivity but depend on the early genes for their even if steroid is absent [84].
full activation [74]. For imaginal tissues, therefore, ecdysteroids mainly provide a
Output genes, such as those for cuticle proteins, are targets of permissive signal that allows ongoing programs of tissue
these transcription factors, as well the ecdysteroids (for example patterning and cell specification to progress. These cells express
[75–77]). A full picture of how all of the members of the cascade EcR-A, which lacks an amino-terminal activation domain. As
act to direct the complex program of cuticle assembly has yet to suggested in Figure 3C, the action of EcR-A likely emphasizes
be made. Genes like DHR3 feed forward to activate late genes the function of EcR–USP to support local chromatin remodeling,
like b-Ftz-F1, expression of which also requires the removal of thereby providing access for other developmental control path-
20E [78]. b-Ftz-F1 is involved in some later cuticle gene induction ways to drive the premetamorphic changes [85]. Indeed, a major
(for example [76]), but it also induces production of the receptor function of ecdysteroids in imaginal tissues is to remove the
for ecdysis triggering hormone, a key hormone that controls the repression imposed by its own receptor system. Consequently,
shedding of the cuticle [79]. Hence, maintaining elevated 20E mutations that block ecdysone production result in much more
levels at the end of the molt prevents b-Ftz-F1 expression and severe developmental deficits than do mutants that remove
the shedding of the old cuticle. either EcR or USP (for example [55]). Interestingly, a similar phe-
Ecdysteroids and Metamorphosis nomenon is seen in some tissues in Xenopus tadpoles in which
The ecdysteroids also drive metamorphic development in a the removal of TRa allows some tissues to complete metamor-
manner analogous to that of thyroid hormone and amphibian phosis in the absence of thyroid hormone [86].
metamorphosis [49,50]. Amphibian tadpoles begin in a pre-
metamorphosis phase characterized by nutrient-dependent Juvenile Hormone and Control of the Nature of the Molt
growth of the tadpole, including its limb buds. The secretion of The JHs (Figure 5A) control the entry into metamorphosis and
thyroxine (T4) by the thyroid gland causes prometamorphosis allow the ecdysteroid pulses to evoke different types of molts
characterized by TRa expression, accelerated growth and [87]. The JHs are a family of sesquiterpenes, the most common
patterning of limb buds, and the induction of TRb. The devel- being JH-III, epoxymethylfarnesoate. Depending on species or
oping tissues subsequently acquire deiodinase activity permit- stage, the structure of JH varies in the number of the epoxide
ting the local conversion of T4 to triiodothyronine (T3) [80]. groups and methyl versus ethyl side branches. JH is synthesized
Only T3 is capable of causing metamorphic climax and it acts by paired cephalic glands, the corpora allata. It has a wide variety
through TRb to evoke the dramatic events of tail loss and the of functions, the chief of which being the control over the pro-
eruption of the forelimbs. gression through the different life stages during growth and
For insects, the period from hatching until the start of the last metamorphosis [87]. It may also regulate aspects of reproduc-
larval stage is analogous to amphibian premetamorphosis — tion [88,89], as well as controlling pigment polymorphisms, caste
both larval tissues and imaginal primordia grow in a nutrient- determination in social insects, behavioral control in worker hon-
dependent fashion through this period. This ‘premetamorpho- eybees, and some seasonal polymorphisms [63]. I will confine
sis’ status is largely maintained by the presence of JH (see my discussion of JH to its actions on growth and metamor-
below) and restricts ecdysteroid action to that of molting. phosis.
When the growth of the larva or nymph exceeds its threshold JH is a lipid-soluble hormone and readily penetrates the
size for metamorphosis [81], its next molt will be to the last larval cuticle. Low doses applied topically can locally suppress meta-
or nymphal instar. In the last larval instar of the Holometabola, morphosis in the underlying epidermis, while the rest of the in-
tissues undergo a switch called pupal commitment [64]. As in sect moves ahead in its life history. This was strikingly illustrated
prometamorphosis in amphibians, during pupal commitment by the insect physiologist V.B. Wigglesworth, who wrote his
gene programs switch from simple growth to preparing tissues initials in nymphal cuticle on the back of metamorphosing
for metamorphosis. These changes are seen in larval tissues Rhodnius by the judicious application of a JH-containing extract
(for example [82]) and in imaginal discs and primordia that [90]. The ability of JH and its mimics to act through surface

Current Biology 29, R1252–R1268, December 2, 2019 R1259


Current Biology

Review

A B Figure 5. The juvenile hormone (JH) system


Brain Hemimetabolous Holometabolous
of insects.
(A) JH is synthesized and released from the paired

JH titer
CA JH-III Corpora Allata (CA) situated posterior to the brain.
JH-III is the most widely occurring JH. (B) Typical
JH titers for a hemimetabolous and a holometab-
N-4 N-5 (last) Adult L-4 L-5 (last) Pupa Adult olous insects based on titers from locusts [138],
and the moth Manduca sexta [139,140]. Shaded
regions are periods of the molt. (C) The pathway
C JH D E JH Met Kr-h1 for the developmental actions of JH [after 94]. JH
X X binds to its receptor, Met, which is bound to
cytoplasmic hsp83 in the absence of ligand. JH-
JH
hsp83

met Met then translocates to the nucleus where it


dimerizes with Taiman (Tai) and binds to JH
response elements (JHRE). The main JH target for
developmental functions is the Krüppel-homolog
JH Kr-h1 1 (Kr-h1) transcription factor. (D) Pupae of the
tai met beetle Tribolium castaneum resulting from larvae
JHRE injected with double-stranded RNA in the 4th larval
egfp dsRNA TcMet (RNAi) N5 Precocious adults Adult instar. Larvae receiving a control construct (egfp)
Current Biology undergo normal subsequent larval molts and
make normal sized pupae; those receiving Met
RNAi, mimicked the effects of JH removal and
precociously formed miniature pupae. From [1], copyright (2007) National Academy of Sciences, U.S.A. (E) Examples from Pyrrhocoris apterus showing that
RNAi-mediated knockdown of components in the JH response pathway in the 4 nymphal instar lead to formation of precocious adults rather than 5th instar
th

nymphs. Images from [97].

contact has made this hormone a major target for developing genetics [98]. Furthermore, Met was shown to bind JH and JH
specific agents for insect control [67]. mimics at physiological concentrations, and the interaction of
The general pattern of the JH titer is fairly consistent during the ligand with residues in the ligand-binding pocket has been
embryonic and postembryonic life. During embryogenesis, JH resolved for Gce [99].
is sometimes present in the newly laid egg but it is rapidly metab- Met (or Gce) heterodimerizes with Taiman, another bHLH-PAS
olized and is absent through early to mid-embryogenesis, but family member also known as Steroid Response Coactivator
then appears after dorsal closure as the tissues of the fully- (SRC) or b-Ftz-F1 Interacting Steroid Coactivator (FISC) [94]
grown embryo mature and the last embryonic cuticle is depos- (Figure 5C). This interaction is intriguing since Taiman also par-
ited [91–93]. The subsequent growth of both larvae and nymphs ticipates in ecdysone action as part of a coactivator complex
occurs in an environment of high circulating JH (Figure 5B). In with EcR–USP. The main target of JH–Met is the gene Krüppel
hemimetabolous insects, JH levels are high through the penulti- homolog 1 (Kr-h1), which encodes a C2H2 zinc finger transcrip-
mate nymphal stage, but then drop during the molt to the last tion factor. In many species, the removal of Kr-h1 mimics both
nymphal stage. JH is then absent through the molt to the adult. the loss of JH and the loss of Met (Figure 5E). JH is also thought
In holometabolous insects, the pattern is similar in that JH is pre- to have an as-yet unidentified membrane receptor that may be
sent as larvae progress through their instars, but then disappears especially important for some of the reproductive functions of
early in the last larval instar to set the stage for the formation of JH [94]. If this membrane receptor has a role in metamorphosis,
the pupa. JH then transiently reappears during the prepupal ec- it is likely supportive to Met/Gce, as removal of the latter fully
dysteroid pulse that directs the molt to the pupa, but then disap- mimics removal of JH.
pears to allow adult differentiation. The Actions of JH
The JH receptor The embryonic actions of JH are poorly understood. JH typically
The developmental effects of JH functions are mediated through appears during embryogenesis around the time of dorsal closure
an intracellular receptor from the basic helix-loop-helix Per/Arnt/ as the embryo shifts from growth and patterning to differentiating
Sim (bHLH-PAS) transcription factor family [94]. It is related to the tissues of the first instar nymph or larva [26]. Embryos vary
the vertebrate aryl-hydrocarbon receptor, but is the only family greatly, though, in their sensitivity to exogenous JH given before
member co-opted to serve as a hormone receptor. It is encoded this time. They appear to be insensitive to JH through segmenta-
by the Methoprene tolerant (Met) gene, which was first identified tion and establishment of limb buds and other primordia, a
in Drosophila in a screen for mutants resistant to the JH mimic period that is concluded by the first embryonic molt. In embryos
methoprene [95]. Establishing that Met was a JH receptor was from more basal groups, such as grasshoppers and crickets, JH
complicated in Drosophila because this fly has a second recep- treatment after this period arrests morphogenesis and the next
tor gene, germ cell expressed (gce), which gave rise to Met by ecdysone pulse induces premature tissue differentiation
gene duplication [96]. Demonstration that Met, indeed, mediates [26,33]. This action of JH in suppressing tissue patterning is in
the anti-metamorphic actions of JH came from insects that have line with its morphostatic action on the imaginal primordia
only one Met gene, such as the beetle Tribolium and the bug described below. In contrast to what is seen with basal hemime-
Pyrrhocoris (Figure 5D,E). In both, Met knock-down by RNA tabolous species, on holometabolous embryos JH treatment has
interference (RNAi) caused both premature metamorphosis comparatively mild effects, primarily being confined to suppres-
and the loss of sensitivity to exogenous JH [1,97]. Definitive proof sion of morphogenetic movements (blastokinesis) within the egg
that Met and Gce are JH receptors then came using Drosophila [100]. Indeed, most of the morphogenetic processes that are

R1260 Current Biology 29, R1252–R1268, December 2, 2019


Current Biology

Review

sensitive to JH treatment in grasshopper embryos have been Variations in Requirements for JH


shifted into the postembryonic realm in the Holometabola, where Insect larvae vary dramatically in how they respond to implanta-
they still display their JH sensitivity. tion of active corpora allata or treatment with exogenous JH.
One embryonic action of JH is opposite to what was ex- Many beetle larvae and caterpillars respond by undergoing addi-
pected from its postembryonic effects. As described below, tional larval molts and becoming giant larvae. By contrast, the
JH is generally considered to have a ‘status quo’ effect on same treatments will not cause larvae of Diptera to make an ex-
molting, causing the insect to produce the same type of cuticle tra larval instar. In the former species, the number of larval instars
that it had had before. In embryonic crickets and grasshoppers, is somewhat plastic and can be modified by nutritional manipu-
though, JH treatment after the E1 molt results in a jump to a lations, while in the latter the larval instar number is fixed. It has
nymphal (E3) molt, rather than repeating the E1 molt [26,33]. become clear recently, though, that even in the species that
The reverse experiment of chemically preventing JH production show classic JH-mediated plasticity, not all of the instars are
in embryos of the cockroach, Nauphoeta cinerea, had the sensitive to JH [107–109].
opposite effect of blocking the appearance of nymphal features Classic experiments in a blood-sucking bug, Rhodnius
[101]. In Blattella germanica, the use of maternal RNAi to sup- prolixus, and the greater wax moth, Galleria mellonella, showed
press the production of JH or its receptor produced a variety of that a first instar nymph or the epidermis from first instar larvae
phenotypes, one of which being an arrest as an apparent pro- could be caused to directly undergo metamorphosis if exposed
nymph [102]. More needs to be done, but in the embryos of to the hormonal environment of the last nymphal or larval stage.
these hemimetabolous insects JH appears to serve to promote These results were the foundation of the idea that the presence
nymph development. of JH was essential for every larval or nymphal molt. Recently,
JH has two main categories of action during postembryonic though, it was shown that larvae of Bombyx mori which lacked
life. Its classic status quo action is linked to the ecdysteroid pulse the JH receptor or the enzymes to make JH nevertheless pro-
that causes a molt [103]. The presence or absence of JH at the gressed through the first two larval molts and only initiated meta-
beginning of the ecdysteroid pulse affects the ‘commitment’ of morphosis after the 3rd larval instar [107,108]. These studies, and
the cells, thereby determining whether the molt will be a progres- similar ones in the bug Pyrrhocoris apterus [108], show that JH is
sive molt to the next stage or a status quo molt repeating the not required for the first few molts of the juvenile.
same stage [64]. Commitment occurs early in the preparatory These results suggest that a metamorphosis-promoting factor
phase so that required cell divisions and the reprogramming of is released after the first few molts and, once it appears, JH is
response genes occurs before the high levels of 20E cause then needed to suppress metamorphosis and maintain the juve-
cuticle production. A modified relationship is seen during the nile stage. The instar at which this metamorphosis-promoting
larval–pupal transition, in which pupal commitment of larval cells factor first appears may determine whether or not there is JH-
is caused by the small ecdysteroid pulse that terminates feeding controlled plasticity in the number of instars. For example, one
and induces a search for a pupation site. The pupal molt itself is might speculate that in Drosophila such a factor does not appear
delayed for hours to days until the occurrence of the large prepu- until after the larva has crossed the threshold size for metamor-
pal peak of ecdysteroids. phosis and, hence, the last larval stage has already been estab-
A second class of JH action has to do with controlling the form lished. There is an increasing list of peptides that influence insect
that the insect will take at the next molt, the morphostatic action growth [110], such as ones from brain, from the fat body, and
of JH [104]. As detailed above, the larva carries imaginal from growing imaginal discs. How these might relate to meta-
primordia either as invaginated imaginal discs or as regions of morphosis-promoting factors is not clear but will likely provide
diploid cells that have the capacity to partially dedifferentiate an intriguing story for the future.
and undergo pupal morphogenesis. The tonic presence of JH The progression through larval instars was thought to be very
in the intermolt periods in preterminal larval instars acts to pre- different in nematodes as compared to insects, as in the former
vent these discs and primordia from initiating premature the progression is controlled by a hard-wired genetic circuit of
morphogenesis [41]. heterochronic genes [111], while insects were thought to use
In hemimetabolous insects, JH disappears late in the molt to hormonal switches to progress through their series. It now ap-
the last nymphal stage to allow the formation of the adult pears that the early set of instars in insects are likely also
(Figure 5B). Holometabolous insects also show the clearing of ‘hard-wired’ while the last few instars, in which most of the
JH from the circulation of last stage larvae. In the latter, though, growth occurs, have a hormone-regulated plasticity to deal
JH reappears during the prepupal ecdysteroid pulse. The reap- with nutritional challenges. We need to reexamine if some of
pearance of JH has two functions. In some insects it ensures the heterochronic genes of nematodes might also function in in-
that imaginal tissues will undergo a pupal molt rather than sects to regulate the early, ‘hard-wired’ portion of their larval/
skipping directly to the adult [103]. A second function relates nymphal series.
to the fact that some of the proliferation and cell patterning for
adult structures begins in the late larva and extends through Stage-Specification Genes and Life History Strategies
the pupal molt into the early phases of adult development, An important advance over the past decade has been the appre-
when they can be turned off by high levels of 20E. Similar high ciation of the role of stage-specification genes. By analogy to the
20E levels occur during the prepupal peak and the presence of role of the homeotic genes that establish differences along the
JH prevents these from permanently switching off proliferation spatial axis of the insect, the stage-specification genes establish
and patterning [105]. JH appears to play a similar role during pu- stage differences along its temporal axis. They are the embodi-
pariation in Drosophila [106]. ment of ‘commitment’ and they determine the character of the

Current Biology 29, R1252–R1268, December 2, 2019 R1261


Current Biology

Review

Ai ii B
Br
+ +

E93
+
L
Br/E93
P
E93
A L
X
Br/E93
L-A mosaic
JH
Kr-h1
iii iv Ecdysteroids Low 20E
MIFs?

D JH 20E
P2
C L8 L9 L10
JH/Met
L
X
Br/E93
L L L
X
P E93 A
Kr-h1 Early genes
L (High thresh)
C Br PN
L1 Early/late
N1 L1
E93 Broad Early genes
(Low thresh)
JH Late
Kr-h1
LF
LF Larval
Pupal
NF Pu Adult
Pu
Effector gene sets
E93
A A

Current Biology

Figure 6. Role of the stage-specification genes in directing insect life histories.


(A) RNAi-based knock-down of broad or E93 in Tribolium castaneum alters progression through the life cycle. (i) Relationship of broad (Br) and E93 to the normal
progression from larva (L) to pupa (P) to adult (A). (ii) Removal of Broad causes the last instar larva to molt to a larval-adult mosaic rather than to a pupa. Images in
(Ai) and (Aii) republished with permission of The Company of Biologists, from [119] ª 2008. (iii) Knockdown of E93 in the larva results in multiple larval molts rather
that forming a pupa (C). Adapted from [125]. (iv) Knockdown of E93 in the pupa results in a second pupa (P2) rather than an adult (from [3]). (B) Levels of expression
of broad (Br), E93, and Krüppel homolog 1 (Kr-h1) in Tribolium castaneum showing how their levels relate to JH and molting peaks of ecdysteroids (red and black
bars) and to interactions amongst the various genes (levels from [125,128]). Other metamorphosis-inducing factors (MIFs) may be involved in broad and E93
induction in some last larval instar tissues. (C) Summary of the shift in expression of Kr-h1, broad, and E93 in hemimetabolous (such as cockroach) and basal
(Tribolium) and derived (higher flies) holometabolous insects. Broad expression is a feature of nymphal (N) and pupal (Pu) stages and is maintained by JH acting
through Kr-h1. The preceding pronymph (PN) and larval (L) stages are characterized by little or no broad expression. E93 determines the adult stage in all cases
and serves as the gateway for metamorphosis for hemimetabolous species and basal holometabolans, but its gateway function is taken over by broad in derived
Holometabola. (D) Intersection of the pathways downstream of JH and of ecdysteroids to determine the stage-specific nature of the larval, pupal and adult molts.
The boxed genes are components of the Ashburner cascade that are common to all molts.

ensuing molt. These genes, summarized in Figure 6, typically domain but vary in their paired C2H2 zinc fingers at their
include Kr-h1, broad and Ecdysone-inducible protein 93F carboxyl terminus [114]. The BTB domain can interact with
(E93). All were initially identified in Drosophila and associated both co-repressors and co-activators as well as supporting
with ecdysteroid action during metamorphosis. The function of chromatin remodeling [115]. The precise interactions that
broad to establish the pupal stage was initially demonstrated in Broad uses, though, are not well understood. Broad protein ap-
Drosophila [112,113], but the scope of the functions of E93 and pears as larval tissues become committed to pupal develop-
Kr-h1 was only revealed by work on insects with less derived ment [116] and Drosophila mutants lacking all broad isoforms
development. stay as last stage larvae and do not initiate their premetamor-
In holometabolous insects, Kr-h1, broad, and E93 direct the phic programs [112]. Experiments in Drosophila show that
genetic programs associated with the larva, pupa, and adult, Broad suppresses both larval- and adult-related programs
respectively. Broad (also called Broad Complex, Br-C) was [113,117] while activating genes specific to the pupal stage.
the first to be discovered and its action is best known. It be- This dual action of Broad is strikingly evident in Tribolium, in
longs to the Bric-a-brac-Tramtrack-Broad zinc finger (BTB- which Broad knock-down by RNAi results in the last stage larva
ZF) family of transcription factors. Depending on species, there becoming a larva-adult mosaic rather than a pupa (Figure 6A)
may be four to six main isoforms that share a common BTB [118,119].

R1262 Current Biology 29, R1252–R1268, December 2, 2019


Current Biology

Review

In hemimetabolous insects, broad is expressed during the act in a ‘nymphal’ fashion, because JH now maintains broad
nymphal instars but then disappears in the last instar when the expression. Indeed, pupae treated with JH maintain broad
nymph transforms into the adult [120–122]. This expression expression and undertake a second pupal molt rather than
pattern would seem to support the proposal that the nymphal becoming an adult [113]. Depending on species, Broad and
stage corresponds to the holometabolous pupa, but RNAi- E93 may inhibit each other [113,130]
based knockdown of broad in early stage nymphs of the bugs
Oncopeltus fasciatus and Pyrrhocoris apterus and the cock- Evolutionary Implications of the Stage-Specification
roach Blattella germanica did not cause precocious metamor- Genes and their Control
phosis. It did, though, prevent the nymphal wing pads from There has been a continuing controversy over how the nymphal
shifting to a premetamorphic pattern of enhanced growth and adult stages of hemimetabolous forms evolved into the
[97,120,122], suggesting an ancestral role for broad in controlling larval, pupal and adult stages of the Holometabola. Broad and
wing growth and morphogenesis. However, recent work on the E93 provide important insights into this controversy but, on first
cricket, Gryllus bimaculatus, showed that knock-down of Broad sight, they lead to somewhat conflicting conclusions. The broad
does, indeed, cause precocious metamorphosis [123]. The dif- data support the homology of the nymph to the pupa. The hemi-
ferences between the effects in crickets versus the other species metabolous pronymph and the holometabolous larva express
may reflect differences in the role of broad or differences in the relatively little broad, but this gene is then expressed prominently
effectiveness of RNAi-mediated gene knockdown. The nymphal in the nymph and the pupa, respectively. The important issue
wing pads have the highest levels of broad expression [122] and discussed above is whether broad acts as a stage-specification
an incomplete knock-down in some species might affect the gene for the nymph, as it does for the pupa. Resolution of this
wings but not the rest of the animal. Further work is obviously point is needed to establish broad’s ancestral function and will
needed to establish whether broad is the stage-specification provide a clearer insight into the origin of the pupal stage.
gene for the nymph. The role of E93 has presented a somewhat different picture. It
E93 encodes a helix-turn-helix transcription factor. It was orig- is clear that E93 determines the adult stage in both hemimetab-
inally studied in Drosophila because it was a stage-specific early olous and holometabolous insects. In hemimetabolous nymphs,
puff that first appeared in the late prepupal salivary gland [124]. JH, acting through Kr-h1, suppresses E93 and thereby maintains
Gene knockdown experiments in both hemimetabolous and the nymphal state. In the absence of Kr-h1, E93 is expressed and
holometabolous insects showed that E93 specifies the adult causes adult differentiation. This has been referred to as the
stage [3]. In the absence of E93, nymphal cockroaches stay as MEKRE93 pathway [24], with E93 providing the gateway into
permanent nymphs. Similarly, in the beetle Tribolium the adult metamorphosis. This is also the case for the beetle Tribolium
form is not made if E93 expression is suppressed. Suppression castaneum. The drop in JH and Kr-h1 levels in the last stage larva
of E93 in the last instar larva results in continued larval molting allows E93 expression, but the result is a pupa rather than an
[125], while knock-down in the pupa results in the molt to another adult. Referenced to the initial induction of E93, the beetle shows
pupal stage [3] (Figure 6A). E93 functions as a pioneer protein, a two-molt progression of larva to pupa and then pupa to adult
opening up sites throughout the chromatin to provide accessi- rather than only a single molt to the adult. This is consistent
bility to the gene sets needed to make the adult [126]. It also re- with the idea that the pupa arose from the terminal molt being
presses broad expression by closing the chromatin around the divided into two parts to accommodate the transition from larva
gene [126]. to adult [23]. In the beetle, though, E93 expression is transient
Kr-h1 is often also considered as the stage-specification gene because JH and Kr-h1 reappear during the larval–pupal transi-
for the nymphal and larval states and, therefore, provides a tion to suppress E93 and allow broad to be expressed to cause
reason for equating nymphs and larvae [24,97,127]. Its involve- pupal differentiation. The disappearance of JH after pupal
ment in larval and nymphal molting, though, is in the context of ecdysis then allows E93 expression, the suppression of broad,
its function in mediating the action of JH [128,129]. The early and adult differentiation.
larval molts that do not require JH also do not require Kr-h1 Although E93 expression provides the entry into metamor-
[108], so Kr-h1 is not needed for larval molting per se. Also, phosis in Tribolium, it apparently does not serve this role in
Kr-h1 later switches to controlling the formation of the pupa in more derived Holometabola, such as Drosophila. E93 still deter-
the cases in which JH reappears in the prepupa to prevent pre- mines the adult stage but it is not expressed until late in the pre-
cocious adult formation (for example [130]). pupal period [131]. Besides specifying the pupal stage in flies,
Kr-h1, broad, and E93 expression are all sensitive to hormonal broad has taken over the role of metamorphic gatekeeper as
signals and they also interact with each other (Figure 6B). As shown by broad null mutants arresting as premetamorphic
noted above, Kr-h1 is the principal target of JH and levels of larvae [112].
Kr-h1 protein are often used as a proxy for the JH titer. Broad This shift from E93 to broad in controlling the entry into meta-
and E93 expression are induced by ecdysteroids but they may morphosis may explain the variation seen within the Holometa-
also be induced by other factors that appear during the last larval bola in the need for JH/Met/Kr-h1 signaling to make a pupa.
instar (for example [41]). Kr-h1 suppresses E93 expression in While lack of Kr-h1 expression during the larval–pupal transition
both cockroaches [24] and Tribolium [130]. The relationship of of Tribolium results in imaginal primordia undergoing adult,
JH/Kr-h1 to broad expression, though, is complex. JH maintains rather than pupal, differentiation, a similar lack of JH in many
broad expression in nymphs [120, 122], but in holometabolous lepidopteran larvae, such as Bombyx mori, or in Drosophila,
larvae JH inhibits the appearance of Broad [116]. But once larval does not. In Tribolium, the use of E93 to initiate metamorphosis
cells become pupally committed by broad induction, they now means that it must then be suppressed by the JH/Met/Kr-h1

Current Biology 29, R1252–R1268, December 2, 2019 R1263


Current Biology

Review

pathway to allow broad expression and pupal development Conclusions


[125]. With the delay of E93 expression in Drosophila, though, The lives of insects are dominated by molts which segment their
JH is no longer needed during the larval–pupal transition to sup- life history into discrete instars that allow growth and changes in
press precocious adult differentiation. JH is still present at that morphology. Each molt is initiated and coordinated by a large
period in flies, though, because of its second function of prevent- pulse of ecdysteroids that act through a conserved genetic cir-
ing the high 20E levels of the pupal molt from terminating meta- cuit, the Ashburner cascade, that translates features of the ste-
morphic patterning programs [105,106] roid pulse into the synthesis of a new cuticle and the shedding
of the old one (Figure 6D). The character of the molt, though, is
Further Insights from Derived Life Cycles determined by stage-specification genes that select the effector
Kr-h1, broad and E93 are also key genes involved in the highly gene sets that are acted on by the cascade. In the Holometabola,
derived life cycles seen in some insects. Within the hemimetab- the specification genes E93, and likely also broad, function as
olous orders, a holometabolous-like life history has evolved in pioneer proteins that open up large stretches of chromatin allow-
thrips (order Thysanoptera) and in some Hemiptera: the scale in- ing access to the transcription factors needed for adult and
sects (Coccidae), mealybugs (Pseudococcidae), and whiteflies pupal differentiation, respectively. JH, acting through the tran-
(Aleyrodidae). Thrips have two larval stages that feed, followed scription factor Kr-h1, exerts the major control over the stage-
by two to three non-feeding pupal stages before making the specification genes. How the developmental hormones and
adult. Broad expression is low in the first larval stage, but then stage-specification genes relate to the evolution of the nymph
rises in the second larval instar in preparation for the production and adult pattern of hemimetabolous insects into the larva/
of the propupa (the first pupal instar) and continues during the pupa/adult pattern of the Holometabola has been a persisting
production of the pupa [7]. In the hemipterans, only the males issue of controversy.
show a complex metamorphosis. As in thrips, in the Japanese The evolution of the homoletabolous larva involved
mealy bug, Planococcus kraunhiae, broad transcripts appear fundamental alterations in embryogenesis. In hemimetabolous
as the feeding ‘larva’ molts to a prepupa and then to a pupa orders, embryogenesis typically generates a miniature approxi-
[132]. In both thrips [133] and mealy bugs [6], the expression of mation of the adult form. In the Holometabola, by contrast,
E93 accompanies the expression of broad as they enter the pu- embryogenesis is altered to produce a larva whose form has
pal series, but then E93 carries on alone in making the adult. little relationship to that of the adult. Ancestral programs of
This parallel evolution of a holometabolous-like life cycle in the growth and patterning are truncated, and the resulting struc-
Neometabola, a group of insects that are clearly members of the tures adapted for use in the new larval form. This truncation
hemimetabolous clade, may provide valuable insights into two also resulted in persisting primordia that undergo deferred
aspects of the changes that gave rise to the Holometabola. morphogenesis at the end of larval growth to make the first
Firstly, their metamorphosis from their larval to their adult stages approximation of the adult in the form of the pupa. The phases
does not involve one pupa-like instar, but rather two to three of embryonic growth and patterning altered to make the larva
such instars. This differs from the idea that the terminal nymphal were those that occur during a cryptic embryonic stage, the
instar became specialized as the homometabolous pupa (for pronymph, that precedes the formation of the nymph and the
example [22]). The neometabolous strategy is more like our pro- normal appearance of JH. Given the sensitivity of the pronymph
posal that the conversion of the pronymph into a feeding larval stage to precocious JH exposure, a heterochronic advance-
stage resulted in selection pressure to concentrate feeding activ- ment of JH into this earlier stage may have caused the
ity in the larva and make the nymph a transition stage [26]. This patterning arrest needed to form the larva. The pronymph
transformation was then followed by a reduction in the number shows low levels of broad compared to the nymph and this
of nymphal (pupal) instars to a single pupal instar evident in Hol- feature of little or no broad expression in the larva is consistent
ometabola today. Thrips may also be in the process of such a with its origin from the pronymph stage.
reduction with one suborder (the Tubulifera) having three pupa- Although there is conflicting evidence as to whether or not
like instars while the other (Terebrantia) has reduced the pupa broad is really a nymphal stage-specification gene, its expres-
to two instars. sion is clearly a feature of nymphs (Figure 6C). JH, through
Although the males of the Hemiptera that show a neometabo- Kr-h1, maintains broad expression through the nymphal instars.
lous lifestyle make a mobile, winged male, the females are neo- A similar relationship of Broad to JH is also evident in the holo-
tenous and remain sessile and larva/nymph-like. Typically, these metabolous pupa. Once larval cells become pupally committed
females reproduce asexually for part of the year and only start and begin expressing broad, JH can no longer turn off that
producing males under times of stress. Expression of E93 is expression and, indeed, JH is necessary for broad expression
very low in such females, associated with their failure to assume during the pupal molt to prevent precocious adult differentiation.
an adult form [6]. Moreover, treatment of pupae with JH supports a reinduction of
Within the Holometabola, neoteny has also been developed in broad resulting in second pupa rather than an adult. What we do
the Strepsiptera, an order of parasitic insects that are related to not understand, though, is how JH suppresses broad expression
beetles and are parasitic on wasps and Hemiptera. The female in larval stages while it promotes broad expression in nymphs.
remains larviform and never leaves the host while the male The answer may lie in the control of broad expression in the pro-
undergoes a normal metamorphosis. The male shows the ex- nymph, a relationship that has not been examined.
pected expression of broad and E93 during its metamorphic pro- E93 determines the adult stage in both hemimetabolous and
gression while the female expresses little, if any, of these two holometabolous insects, but its relationship to both broad and
transcription factors [4,5]. JH (via Kr-h1) is complex and different features support

R1264 Current Biology 29, R1252–R1268, December 2, 2019


Current Biology

Review

alternative views to the origin of the pupa. The nymph-adult tran- 7. Minakuchi, C., Tanaka, M., Miura, K., and Tanaka, T. (2011). Develop-
mental profile and hormonal regulation of the transcription factors broad
sition is simple: JH, acting through Kr-h1, suppresses E93 and and Krüppel homolog 1 in hemimetabolous thrips. Insect Biochem. Mol.
maintains broad; the decline in JH in the last nymphal instar al- Biol. 41, 125–134.
lows E93 expression, thereby both suppressing broad and
8. Nijhout, H.F., and Callier, V. (2014). Developmental mechanisms of body
causing adult differentiation. This relationship is somewhat size and wing-body scaling in insects. Annu. Rev. Entomol. 60, 141–156.
modified in basal holometabolans, as represented by Tribolium.
9. Chafino, S., Ureña, E., Casanova, J., Casacuberta, E., Franch-Marro, X.,
Here the suppression of E93 by JH–Kr-h1 in the larva also con- and Martı́n, D. (2019). Upregulation of E93 gene expression acts as the
trols the entry to metamorphosis and this relationship is taken trigger for metamorphosis independently of the threshold size in the bee-
as support that the larval–pupal transition corresponds to the tle Tribolium castaneum. Cell Rep. 27, 1039–1049.
nymph–adult transition [24,97]. However, the repression of E93 10. Rousset, A., and Bitsch, C. (1993). Comparison between endogenous
by JH–Kr-h1 is seen again in the following molt of the pupa to and exogenous yolk proteins along an ovarian cycle in the firebrat
Thermobia domestica (Insecta.Thysanura). Comp. Biochem. Physiol.
the adult. The pupa–adult molt is more like the last nymphal 104B, 33–44.
molt, because Broad is now present and JH–Kr-h1 functions in
the pupa to maintain broad and inhibit E93. Including Broad in 11. Kukalova-Peck, J. (1991). Fossil history and the evolution of hexapod
structures. In The Insects of Australia, Vol. 1 (Carlton: Melbourne Univer-
the relationship indicates that the pupa–adult molt corresponds sity Press), 141–17.
to the nymph–adult molt.
12. Prokop, J., Percharová, M., Nel, A., Hörnschemeyer, T., Krzeminska, E.,
To finally interpret the relationships that are seen at metamor- Krzeminski, W., and Engel, M.S. (2017). Paleozoic nymphal wing pads
phosis, though, we need more information on the hormonal regu- support dual model of insect wing origins. Curr. Biol. 27, 263–269.
lation of the embryonic molts of more basal species. Hormonal 13. Belles, X. (2019). The innovation of the final moult and the origin of insect
relationships have continued to change after the evolution of metamorphosis. Philos. Trans. R. Soc. Lond. B 374, 20180415.
complete metamorphosis, as seen in highly derived groups
14. Yang, A.S. (2001). Modularity, evolvability, and adaptive radiations: a
such as the higher Diptera (Figure 6C). In these insects the entry comparison of hemi- and holometabolous insects. Evol. Dev. 3, 59–72.
to metamorphosis has now been given over to broad and E93
15. Rainford, J.L., Hofreiter, M., Nicholson, D.B., and Mayhew, P.J. (2014).
expression is delayed until the end of the pupal stage at the start Phylogenetic distribution of extant richness suggests metamorphosis is
of adult differentiation. This shift has diminished the role of the a key innovation driving diversification in insects. PLoS One 9, e109085.
prepupal JH peak, and simplified the endocrine regulation over
16. Nel, A., Roques, P., Nel, P., Prokin, A.A., Bourgoin, T., et al. (2013). The
metamorphosis. earliest known holometabolous insects. Nature 503, 257–261.

17. Sehnal, F., Svácha, P., and Zrzavy, J. (1996). Evolution of insect meta-
ACKNOWLEDGEMENTS morphosis. In Metamorphosis: Postembryonic Reprogramming of
Gene Expression in Amphibian and Insect Cells, L.I. Gilbert, J.R. Tat,
I am grateful to Lynn Riddiford for many discussions during the preparation of and B.G. Atkinson, eds. (San Diego, CA: Academic Press), pp. 3–58.
this article and for critical readings of the manuscript. Important feedback on
18. Erezyilmaz, D.F. (2006). Imperfect eggs and oviform nymphs: a history of
the manuscript was also provided by Marek Jindra, Joel Kingsolver, Stuart ideas about the origins of insect metamorphosis. Integr. Comp. Biol. 46,
Reynolds and anonymous reviewers. I thank Marek Jindra for supplying the 795–807.
images for Figure 5D,E. Many of the ideas arose during the course of research
funded by National Science Foundation, the National Institutes of Health, and 19. Reynolds, S.E. (2019). Cooking up the perfect insect: Aristotle’s transfor-
the Howard Hughes Medical Institute. matopnal idea about the complete metamorphosis of insects. Philos.
Trans. R. Soc. Lond. B 374, 20190074.

20. Berlese, A. (1913). Intorno alle metamorfosi degli insetti. Redia 9,


REFERENCES 121–136.

21. Imms, A.D. (1937). Recent Advances in Entomology (Philadelphia, PA:


1. Konopova, B., and Jindra, M. (2007). Juvenile hormone resistance
Blakiston).
gene Methoprene-tolerant controls entry into metamorphosis in
the beetle Tribolium castaneum. Proc. Natl. Acad. Sci. USA 104, 22. Hinton, H.E. (1963). The origin and function of the pupal stage. Proc. R.
10488–10493. Entomol. Soc. London Ser. A 38, 77–85.
2. Charles, J.P., Iwema, T., Epa, V.C., Takaki, K., Rynes, J., and Jindra, M. orie de la nymphe des Insectes Hol-
23. Poyarkoff, E. (1914). Essai d’une the
(2011). Ligand-binding properties of a juvenile hormone receptor, Metho- taboles. Arch. Zool. Exp. Gen. 54, 221–265.
ome
prene-tolerant. Proc. Natl. Acad. Sci. USA 108, 21128–21133.
24. Belles, X., and Santos, C.G. (2014). The MEKRE93 (Methoprene tolerant-
3. Ureña, E., Manjón, C., Franch-Marro, X., and Martı́n, D. (2014). Transcrip- Krüppel homolog 1-E93) pathway in the regulation of insect metamor-
tion factor E93 specifies adult metamorphosis in hemimetabolous and phosis, and the homology of the pupal stage. Insect Biochem. Mol.
holometabolous insects. Proc. Natl. Acad. Sci. USA 111, 7024–7029. Biol. 52, 60–68.
4. Erezyilmaz, D.F., Hayward, A., Huang, Y., Paps, J., Acs, Z., Delgado, 25. Jindra, M. (2019). Where does the pupa come from? The timing of
J.A., Collantes, F., and Kathirithamby, J. (2014). Expression of the pupal juvenile hormone signaling supports homology between stages of hemi-
determinant broad during metamorphosis and neotenic development of metabolous and holometabolous insects. Philos. Trans. R. Soc. B 374,
the strepsipteran Xenos vesparum. PLoS One 9, e93614. 20190064.
5. Chafino, S., López-Escardó, D., Benelli, G., Kovac, H., Casacuberta, E., 26. Truman, J.W., and Riddiford, L.M. (1999). The origins of insect metamor-
Franch-Marro, X., Kathirithamby, J., and Martı́n, D. (2018). Differential phosis. Nature 401, 447–452.
expression of the adult specifier E93 in the strepsipteran Xenos
vesparum Rossi suggests a role in female neoteny. Sci. Rep. 8, 14176. 27. Truman, J.W., and Riddiford, L.M. (2002). Endocrine insights into the
evolution of metamorphosis in insects. Annu. Rev. Entomol. 47, 467–500.
6. Vea, I.M., Tanaka, S., Tsuji, T., Shiotsuki, T., Jouraku, A., and Minakuchi,
C. (2019). E93 expression and links to the juvenile hormone in hemipteran 28. Truman, J.W., and Riddiford, L.M. (2019). The evolution of insect meta-
mealybugs with insights on female neoteny. Insect Biochem. Mol. Biol. morphosis: a developmental and endocrine view. Philos. Trans. R. Soc.
104, 65–72. B 374, 20190070.

Current Biology 29, R1252–R1268, December 2, 2019 R1265


Current Biology

Review
29. Cohen, S.M. (1993). Imaginal disc development. In The Development of 50. Brown, D.D., and Cai, L. (2007). Amphibian metamorphosis. Dev. Biol.
Drosophila melanogaster, Vol. 2, M. Bate, and A. Martinez-Arias, eds. 306, 20–33.
(Plainview, NY: Cold Spring Harbor Press), pp. 747–841.
51. Warren, J.T., and Gilbert, L.I. (1986). Ecdysone metabolism and distribu-
30. Paulus, H.F. (1986). Evolutionswege zum Larvalauge der Insekten-ein tion during the pupal-adult development of Manduca sexta. Insect Bio-
€ren Lateralaugen
Modell für die Entstehung und die Ableitung der ozella chem. 16, 65–82.
der Myriapoda von Fazettenaugen. Zool. Jahrb. Syst. 113, 353–371.
52. Lagueux, M., Hetru, C., Goltzene, F., Kappler, C., and Hoffmann, J.A.
31. Liu, Z., and Friedrich, M. (2004). The Tribolium homologue of glass and (1979). Ecdysone titre and metabolism in relation to cuticulogenesis in
the evolution of insect larval eyes. Dev. Biol. 269, 36–54. embryos of Locusta migratoria. J. Insect Physiol. 25, 709–723.

32. Inoue, Y., Mito, T., Miyawaki, K., Matsushima, K., Shinmyo, Y., Heanue, 53. Rewitz, K.F., O’Connor, M.B., and Gilbert, L.I. (2007). Molecular evolution
T.A., Mardon, G., Ohuchi, H., and Noji, S. (2002). Correlation of expres- of the insect Halloween family of cytochrome P450s: phylogeny, gene or-
sion patterns of homothorax, dachshund, and Distal-less with the proxi- ganization and functional conservation. Insect Biochem. Mol. Biol. 37,
modistal segmentation of the cricket leg bud. Mech. Dev. 113, 141–148. 741–753.

33. Erezyilmaz, D.F., Riddiford, L.M., and Truman, J.W. (2004). Juvenile hor- 54. Laudet, V., and Gronemeyer, H. (2002). The Nuclear Receptors Facts-
mone acts at embryonic molts and induces the nymphal cuticle in the book. (London: Academic Press).
direct-developing cricket. Dev. Genes Evol. 214, 313–323.
55. Koelle, M.R., Talbot, W.S., Segraves, W.A., Bender, M.T., Cherbas, P.,
34. Kojima, T. (2004). The mechanism of Drosophila leg development along and Hogness, D.S. (1991). The Drosophila EcR gene encodes an ecdy-
the proximodistal axis. Dev. Growth Differ. 46, 115–129. sone receptor, a new member of the steroid receptor superfamily. Cell
67, 59–77.
35. Tanaka, K., and Truman, J.W. (2007). Molecular patterning mechanism
underlying metamorphosis of the thoracic leg in Manduca sexta. Dev. 56. Talbot, W.S., Swyryd, E.A., and Hogness, D.S. (1993). Drosophila tissues
Biol. 305, 539–550. with different metamorphic responses to ecdysone express different
ecdysone receptor isoforms. Cell 73, 1323–1337.
36. Kim, C.-W. (1959). The differentiation center inducing development from
larval to adult leg in Pieris brassicae (Lepidoptera). J. Embryol. Exp. Mor- 57. Watanabe, T., Takeuchi, H., and Kubo, T. (2010). Structural diversity and
phol. 7, 572–582. evolution of the N-terminal isoform-specific region of ecdysone receptor-
A abd -B1 isoforms in insects. BMC Evol. Biol. 10, 40.
37. Tanaka, K., and Truman, J.W. (2005). Development of the adult leg
epidermis in Manduca sexta: contribution of different larval cell popula- 58. Hill, R.J., Billas, I.M., Bonneton, F., Graham, L.D., and Lawrence, M.C.
tions. Dev. Genes Evol. 215, 78–89. (2013). Ecdysone receptors: from the Ashburner model to structural
biology. Annu. Rev. Entomol. 58, 251–271.
38. Konopova, B., and Zrzavý, J. (2005). Ultrastructure, development, and
homology of insect embryonic cuticles. J. Morph. 267, 339–362. 59. Riddiford, L.M., Cherbas, P., and Truman, J.W. (2000). Ecdysone recep-
tors and their biological actions. Vitam. Horm. 60, 1–73.
39. Bernays, E.A. (1971). The vermiform larva of Schistocerca gregaria
(Forskål): form and activity (Insects, Orthoptera). Z. Morph. Tiere 70, 60. Sullivan, A.A., and Thummel, C.S. (2003). Temporal profiles of nuclear re-
183–200. ceptor gene expression reveal coordinate transcriptional responses dur-
ing Drosophila development. Mol. Endocrinol. 17, 2125–2137.
40. Svácha, P. (1992). What are and what are not imaginal discs: reevaluation
of some basic concepts (Insecta, Holometabola). Dev. Biol. 154, 61. Andersen, S.O., Højrup, P., and Roepstorff, P. (1995). Insect cuticular
101–117. proteins. Insect Biochem. Mol. Biol. 25, 153–176.

41. Truman, J.W., Hiruma, K., Allee, J.P., MacWhinnie, S.G.B., Champlin, 62. Moussian, B. (2010). Recent advances in understanding mechanisms of
D.T., and Riddiford, L.M. (2006). Juvenile hormone is required to couple insect cuticle differentiation. Insect Biochem. Mol. Biol. 40, 363–375.
imaginal disc formation with nutrition in insects. Science 312, 1385–1388.
63. Nijhout, H.F. (1994). Insect Hormones (Princeton, NJ: Princeton Univer-
42. Mirth, K.C., Truman, J.W., and Riddiford, L.M. (2009). The Ecdysone re- sity Press).
ceptor controls the post-critical weight switch to nutrition-independent
differentiation in Drosophila wing imaginal discs. Development 136, 64. Riddiford, L.M. (1976). Hormonal control of insect epidermal cell commit-
2345–2353. ment in vitro. Nature 259, 115–117.

43. Lafont, R., Dauphin-Villemant, C., Warren, J.T., and Rees, H. (2012). In 65. Williams, C.M. (1968). Ecdysone and ecdysone-analogues: their assay
Ecdysone chemistry and Biochemistry, Insect Endocrinology, and L.I. and action on diapausing pupae of the Cynthia silkworm. Biol. Bull.
Gilbert, eds. (Amsterdam: Elsevier Press), pp. 106–176. 134, 344–355.

44. Krishnakumaran, A., and Schneiderman, H.A. (1970). Control of molting 66. White, B.H., and Ewer, J. (2014). Neural and hormonal control of postec-
in mandibulate and chelicerate arthropods by ecdysones. Biol. Bull. dysial behaviors in insects. Annu. Rev. Entomol. 59, 363–381.
139, 520–538.
67. Dhadiallah, T.S., Carlson, G.R., and Le, D.P. (1998). New insecticides
45. Qu, Z., Kenny, N.J., Lam, H.M., Chan, T.F., Chu, K.H., Bendena, W.G., with ecdysteroidal and juvenile hormone activity. Annu. Rev. Entomol.
Tobe, S.S., and Hui, J.H. (2015). How did arthropod sesquiterpenoids 43, 545–569.
and ecdysteroids arise? Comparison of hormonal pathway genes in
noninsect arthropod genomes. Genome Biol. Evol. 7, 1951–1959. €nderungen in
68. Clever, U., and Karlson, P. (1960). Induktion von puff-vera
den Speicheldrüsenchromosomen von Chironomus tentans durch
46. Smith, S.L., Bollenbacher, W.E., and Gilbert, L.I. (1983). Ecdysone 20- Ecdyson. Exp. Cell Res. 20, 623–626.
monooxygenase activity during larval-pupal development of Manduca
sexta. Mol Cell Endocrinol. 31, 227–251. 69. Ashburner, M., Chihara, C., Meltzer, P., and Richards, G. (1974). Tempo-
ral control of puffing activity in polytene chromosomes. Cold Spring
47. Champlin, D.T., and Truman, J.W. (1998). Ecdysteroids govern two Harb. Symp. Quant. Biol. 38, 655–662.
phases of eye development during metamorphosis of the moth,
Manduca sexta. Development 125, 2009–2018. 70. Hiruma, K., and Riddiford, L.M. (2010). Developmental expression of
mRNA for epidermal and fat body proteins and hormonally regulated
48. Yamanaka, N., Rewitz, K.F., and O’Connor, M.B. (2013). Ecdysone con- transcription factors in the tobacco hornworm, Manduca sexta.
trol of developmental transitions: lessons from Drosophila research. J. Insect Physiol. 56, 1390–1395.
Annu. Rev. Entomol. 58, 497–516.
71. Mane-Padrós, D., Borràs-Castells, F., Belles, X., and Martı́n, D. (2012).
49. Tata, J.R. (2000). Autoinduction of nuclear hormone receptors during Nuclear receptor HR4 plays an essential role in the ecdysteroid-triggered
metamorphosis and its significance. Insect Biochem. Mol. Biol. 30, gene cascade in the development of the hemimetabolous insect Blattella
645–651. germanica. Mol. Cell. Endocrinol. 348, 322–330.

R1266 Current Biology 29, R1252–R1268, December 2, 2019


Current Biology

Review
72. King-Jones, K., and Thummel, C.S. (2005). Nuclear receptors – a 92. Bergot, B.J., Baker, F.C., Cerf, D.C., Jamieson, G., and Schooley, D.A.
perspective from Drosophila. Nat. Rev. Genet. 6, 311–323. (1981). Qualitative and quantitative aspects of juvenile hormone titers in
developing embryos of several insect species: Discovery of a new JH-
73. Karim, F.D., and Thummel, C.C. (1991). Ecdysone coordinates the timing like substance extracted from eggs of Manduca sexta. In Juvenile Hor-
and amounts of E74A and E74B transcription in Drosophila. Genes Dev. mone Biochemistry. Action, Agonism and Antagonism, G.E. Pratt, and
5, 1067–1079. G.T. Brookes, eds. (Amsterdam: Elsevier/North-Holland Biomedical
Press), pp. 33–45.
74. White, K.P., Hurban, P., Watanabe, T., and Hogness, D.S. (1997). Coor-
dination of Drosophila metamorphosis by two ecdysone-induced nuclear 93. Temin, G., Zander, M., and Roussel, J.P. (1986). Physico-chemical
receptors. Science 276, 114–117. (GC-MS) measurements of juvenile hormone III titres during embryogen-
esis of Locusta migratoria. Int. J. Invert. Reprod. 9, 105–112.
75. Charles, J.P. (2010). The regulation of expression of insect cuticle protein s, X., and Shinoda, T. (2015). Molecular basis of juvenile
94. Jindra, M., Belle
genes. Insect Biochem. Mol. Biol. 40, 205–213. hormone signaling. Curr. Opin. Insect Sci. 11, 39–46.
76. Wang, H.B., Moriyama, M., Iwanaga, M., and Kawasaki, H. (2010). Ecdy- 95. Wilson, T.G., and Fabian, J. (1986). A Drosophila melanogaster mutant
sone directly and indirectly regulates a cuticle protein gene, BMWCP10, resistant to a chemical analog of juvenile hormone. Dev. Biol. 118,
in the wing disc of Bombyx mori. Insect Biochem. Mol. Biol. 40, 453–459. 190–201.

77. Cui, H.Y., Lestradet, M., Bruey-Sedano, N., Charles, J.P., and Riddiford, 96. Baumann, A., Fujiwara, Y., and Wilson, T.G. (2010). Evolutionary diver-
L.M. (2009). Elucidation of the regulation of an adult cuticle gene Acp65A gence of the paralogs Methoprene tolerant (Met) and germ cell expressed
by the transcription factor Broad. Insect Mol. Biol. 18, 421–429. (gce) within the genus Drosophila. J. Insect Physiol. 56, 1445–1455.

78. Woodard, C.T., Baehrecke, E.H., and Thummel, C.S. (1994). A molecular 97. Konopova, B., Smykal, V., and Jindra, M. (2011). Common and distinct
mechanism for the stage specificity of the Drosophila prepupal genetic roles of juvenile hormone signaling genes in metamorphosis of holome-
response to ecdysone. Cell 79, 607–615. tabolous and hemimetabolous insects. PLoS One 6, e28728.

79. Cho, K.H., Daubnerová, I., Park, Y., Zitnan, D., and Adams, M.E. (2014). 98. Jindra, M., Uhlirova, M., Charles, J.P., Smykal, V., and Hill, R.J. (2015).
Secretory competence in a gateway endocrine cell conferred by the nu- Genetic evidence for function of the bHLH-PASS protein Gce/Met as a
clear receptor bFTZ-F1 enables stage-specific ecdysone responses juvenile hormone receptor. PLoS Genet. 11, e1005394.
throughout development in Drosophila. Dev Biol. 385, 253–262.
99. Bittova, L., Jedlicka, P., Dracinsky, M., Kirubakaran, P., Vondrasek, J.,
Hanus, R., and Jindra, M. (2019). Exquisite ligand stereoselectivity of a
80. Brown, D.D. (2005). The role of deiodinases in amphibian metamor- Drosophila juvenile hormone receptor contrasts with its broad agonist
phosis. Thyroid 15, 815–821. repertoire. J. Biol. Chem. 294, 410–423.

81. Nijhout, H.F., Riddiford, L.M., Mirth, C., Shingleton, A.W., Suzuki, Y., and 100. Riddiford, L.M., and Williams, C.M. (1967). The effects of juvenile hor-
Callier, V. (2014). The developmental control of size in insects. Wiley In- mone analogues on the embryonic development of silkworms. Proc.
terdiscip. Rev. Dev. Biol. 3, 113–134. Natl. Acad. Sci. USA 57, 595–601.

82. Lepesant, J.A., Levine, M., Garen, A., Lepesant-Kejzlarvoa, J., Rat, L., 101. Brüning, E., and Lanzrein, B. (1987). Function of juvenile hormone III in
and Somme-Martin, G. (1982). Developmentally regulated gene expres- embryonic development of the cockroach, Nauphoeta cinerea. Int. J.
sion in Drosophila larval fat bodies. J. Mol. Appl. Genet. 1, 371–383. Invert. Reprod. 12, 29–44.

83. Cherbas, L., Hu, X., Zhimulev, I., Belyaeva, E., and Cherbas, P. (2003). 102. Fernandez-Nicolas, A., and Belles, X. (2017). Juvenile hormone signaling
EcR isoforms in Drosophila: testing tissue-specific requirements by tar- in short germ-band hemimetabolan embryos. Development 144, 4637–
geted blockade and rescue. Development 130, 271–284. 4644.

, C., Antoniewski, C., and Truman, J.W. (2005). 103. Williams, C.M. (1961). The juvenile hormone. II. Its role in the endocrine
84. Schubiger, M., Carre
control of molting, pupation, and adult development in the Cecropia silk-
Ligand-dependent de-repression via EcR/USP acts as a gate to coordi-
worm. Biol. Bull. 121, 572–585.
nate the differentiation of sensory neurons in the Drosophila wing. Devel-
opment 132, 5239–5248.
104. Truman, J.W., and Riddiford, L.M. (2007). The morphostatic actions of ju-
venile hormone. Insect Biochem. Mol. Biol. I37I, 761–770.
85. Uyehara, C.M., and McKay, D.J. (2019). Direct and widespread role for
the nuclear receptor EcR in mediating the response to ecdysone in 105. Champlin, D.T., Reiss, S.E., and Truman, J.W. (1999). Hormonal control
Drosophila. Proc. Natl Acad. Sci. USA 116, 9893–9902. of ventral diaphragm myogenesis during metamorphosis of the moth
Manduca sexta. Dev. Genes Evol. 209, 265–274.
86. Buchholz, D.R., and Shi, Y.-B. (2018). Dual function model revised by thy-
roid hormone receptor alpha knockout frogs. Gen. Comp. Endocrinol. 106. Riddiford, L.M., Truman, J.W., Mirth, C.K., and Shen, Y.C. (2010). A role
265, 214–218. for juvenile hormone in the prepupal development of Drosophila
melanogaster. Development 137, 1117–1126.
87. Jindra, M., Palli, S.R., and Riddiford, L.M. (2013). The juvenile hormone
signaling pathway in insect development. Annu. Rev. Entomol. 58, 107. Daimon, T., Kozaki, T., Niwa, R., Kobayashi, I., Furuta, K., Namiki, T.,
181–204. Uchino, K., Banno, Y., Katsuma, S., Tamura, T., et al. (2012). Precocious
metamorphosis in the juvenile hormone-deficient mutant of the silkworm,
88. Raikhel, A.S., Brown, M.R., and Belles, X. (2005). Hormonal control of Bombyx mori. PLoS Genet. 8, e1002486.
reproductive processes. In Comprehensive Molecular Insect Science,
Vol 3, L.I. Gilbert, K. Iatrou, and S.S. Gill, eds. (Amsterdam: Elsevier 108. Daimon, T., Uchibori, M., Nakao, H., Sezutsu, H., and Shinoda, T. (2015).
Press), pp. 433–491. Knockout silkworms reveal a dispensable role for juvenile hormones in
holometabolous life cycle. Proc. Natl. Acad. Sci. USA 112, E4226–E4235.
89. Wyatt, G.R., and Davey, K.G. (1996). Cellular and molecular actions of ju- 109. Smykal, V., Daimon, T., Kayukawa, T., Takaki, K., Shinoda, T., and Jin-
venile hormone. II. Roles of juvenile hormone in adult insects. Adv. Insect dra, M. (2015). Importance of juvenile hormone signaling arises with
Physiol. 26, 1–155. competence of insect larvae to metamorphose. Dev. Biol. 390, 221–230.
90. Wigglesworth, V.B. (1958). Some methods for assaying extracts of the ju- 110. Droujinine, I.A., and Perrimon, N. (2016). Interorgan communication path-
venile hormone in insects. J. Insect Physiol. 2, 73–84. ways in physiology: focus on Drosophila. Annu. Rev. Genet. 23, 539–570.

91. Bergot, B.J., Jamieson, G.C., Ratcliff, M.A., and Schooley, D.A. (1980). 111. Faunes, F., and Larrain, J. (2016). Conservation in the involvement of het-
JH Zero: new naturally occurring insect juvenile hormone from devel- erochronic genes and hormones during developmental transitions. Dev.
oping embryos of the tobacco hornworm. Science 210, 336–338. Biol. 416, 3–17.

Current Biology 29, R1252–R1268, December 2, 2019 R1267


Current Biology

Review
112. Kiss, I., Beaton, A.H., Tardiff, J., Fristrom, D., and Fristrom, J.W. (1988). 127. Lozano, J., and Belles, X. (2011). Conserved repressive function of Krüp-
Interactions and developmental effect of mutations in the Broad- pel homolog 1 on insect metamorphosis in hemimetabolous and holome-
Complex of Drosophila melanogaster. Genetics 118, 247–259. tabolous species. Sci. Rep. 1, 163.

113. Zhou, X., and Riddiford, L.M. (2002). Broad specifies pupal development 128. Minakuchi, C., Namiki, T., and Shinoda, T. (2009). Krüppel homolog 1, an
and mediates the ‘status quo’ action of juvenile hormone on the pupal- early juvenile hormone-response gene downstream of Methoprene-
adult transformation in Drosophila and Manduca. Development 129, tolerant, mediates its anti-metamorphic action in the red flour beetle Tri-
2259–2269. bolium castaneum. Dev. Biol. 325, 341–350.
114. Chaharbakhshi, E., and Jemc, J.C. (2016). Broad-complex, tramtrack,
and bric-à-brac (BTB) proteins: critical regulators of development. Gen- 129. Minakuchi, C., Namiki, T., Yoshiyama, M., and Shinoda, T. (2008). RNAi-
esis 54, 505–518. mediated knockdown of juvenile hormone acid O-methyltransferase
gene causes precocious metamorphosis in the red flour beetle Tribolium
, G.G. (1998). Crystal structure of the
115. Ahmad, K.F., Engel, C.K., and Prive castaneum. FEBS J. 275, 2919–2931.
BTB domain of PLZF. Proc. Natl. Acad. Sci. USA 95, 12123–12128.
130. Ureña, E., Chafino, S., Manjón, C., Franch-Marro, X., and Martin, D.
116. Zhou, B., Hiruma, K., Shinoda, T., and Riddiford, L.M. (1998). Juvenile (2016). The occurrence of the holometabolous pupal stage requires the
hormone prevents ecdysteroid-induced expression of Broad Complex interaction between E93, Krüppel-homolog 1 and Broad-Complex.
RNAs in the epidermis of the tobacco hornworm, Manduca sexta. Dev. PLoS Genet. 12, e1006020.
Biol. 203, 233–244.
131. Mou, X., Duncan, D.M., Baehrecke, E.H., and Duncan, I. (2012). Control
117. Riddiford, L.M., Hiruma, K., Zhou, X., and Nelson, C.A. (2003). Insights of target gene specificity during metamorphosis by the steroid response
into the molecular basis of the hormonal control of molting and metamor- gene E93. Proc. Natl. Acad. Sci. USA 109, 2949–2954.
phosis from Manduca sexta and Drosophila melanogaster. Insect Bio-
chem. Mol. Biol. 33, 1327–1338.
132. Vea, I.M., Tanaka, S., Shiotsuki, T., Jouraku, A., Tanaka, T., and Minaku-
118. Konopova, B., and Jindra, M. (2008). Broad-Complex acts downstream chi, C. (2016). Differential juvenile hormone variations in scale insect
of Met in juvenile hormone signaling to coordinate primitive holometabo- extreme sexual dimorphism. PLoS One 11, e0149459.
lan metamorphosis. Development 135, 559–568.
133. Suzuki, Y., Shiotsuki, T., Jouraku, A., Miura, K., and Minakuchi, C. (2019).
119. Suzuki, Y., Truman, J.W., and Riddiford, L.M. (2008). The role of Broad in Personal communication.
the development of Tribolium castaneum: implications for the evolution of
the holometabolous insect pupa. Development 135, 569–577. 134. Wang, Y.-H., Engel, M.S., Rafael, J.A., Wu, H.-Y., Redei, D., Xie, Q.,
Wang, G., Liu, X.-G., and Bu, W.-J. (2016). Fossil record of stem groups
120. Erezyilmaz, D.F., Riddiford, L.M., and Truman, J.W. (2006). The pupal employed in evaluating the chronogram of insects (Arthropoda: Hexa-
specifier broad directs progressive morphogenesis in a direct-devel- poda). Sci. Rep. 6, 38939.
oping insect. Proc. Natl. Acad. Sci. USA 103, 6925–6930.
135. Wipfler, B., Letsch, H., Frandsen, P.B., et al. (2019). Evolutionary history
121. Piulachs, M.-D., Pagone, V., and Belles, X. (2010). Key roles of the Broad-
of Polyneoptera and its implications for our understanding of early
Complex gene in insect embryogenesis. Insect Biochem. Mol. Biol. 40,
winged insects. Proc. Natl. Acad. Sci. USA 116, 3024–3029.
468–475.

122. Huang, J.-H., Lozano, J., and Belles, X. (2013). Broad-complex functions 136. Labandeira, C. C. 1994. A compendium of fossil insect families. Milwau-
in postembryonic development of the cockroach Blattella germanica kee Public Museum Contributions, no. 88: 1–71.
shed new light on the evolution of insect metamorphosis. Biochim. Bio-
phys. Acta 1830, 2178–2187. 137. Bollenbacher, W.E., Smith, S.L., Goodman, W., and Gilbert, L.I. (1981).
Ecdysteroid titer during larval-pupal-adult development of the tobacco
123. Ishimaru, Y., Tomonari, S., Watanabe, T., Noji, S., and Mito, T. (2019). hornworm, Manduca sexta. Gen. Comp. Endocrinol. 44, 302–306.
Regulatory mechanisms underlying the specification of the pupal-homol-
ogous stage in a hemimetabolous insect. Philos. Trans. R. Soc. Lond. B 138. Huibregtse-Minderhoud, L., Van den Hondel-Franken, M.A.M., Van der
374, 20190225. Kerk-Van Hoof, A.C., Biessels, H.W.A., Salemink, C.A., Van der Horst,
D.J., and Beenakkers, A.M.T. (1980). Quantitative determination of the ju-
124. Baehrecke, E.H., and Thummel, C.S. (1995). The Drosophila E93 gene venile hormones in the haemolymph of Locusta migratoria during normal
from the 93F early puff displays stage- and tissue-specific regulation development and after implantation of corpora allata. J. Insect Physiol.
by 20-hydroxyecdysone. Dev. Biol. 171, 85–97. 26, 627–631.
125. Chafino, S., Ureña, E., Casanova, J., Casacuberta, E., Franch-Marro, X.,
and Martin, D. (2019). Upregulation of E93 gene expression acts as the 139. Baker, F.C., Tsai, L.W., Reuter, C.C., and Schooley, D.A. (1987). In vivo
trigger for metamorphosis independently of the threshold size in the bee- fluctuation of JH, JH acid, and ecdysteroid titer, and JH esterase activity
tle Tribolium castaneum. Cell Rep. 27, 1039–1049. during development of fifth stadium Manduca sexta. Insect Biochem. 17,
989–996.
126. Uyehara, C.M., Nystrom, S.L., Niederhuber, M.J., Leatham-Jensen, M.,
Ma, Y., Buttitta, L.A., and McKay, D.J. (2017). Hormone-dependent con- 140. Fain, M.J., and Riddiford, L.M. (1975). Juvenile hormone titers in the he-
trol of developmental timing through regulation of chromatin accessi- molymph during later larval development of the tobacco hornworm,
bility. Genes Dev. 31, 862–875. Manduca sexta (L.). Biol. Bull. 149, 506–521.

R1268 Current Biology 29, R1252–R1268, December 2, 2019

You might also like