You are on page 1of 10

REVIEWS

How flies get their size: genetics


meets physiology
Bruce A. Edgar
Abstract | Body size affects important fitness variables such as mate selection, predation
and tolerance to heat, cold and starvation. It is therefore subject to intense evolutionary
selection. Recent genetic and physiological studies in insects are providing predictions
as to which gene systems are likely to be targeted in selecting for changes in body size.
These studies highlight genes and pathways that also control size in mammals: insects
use insulin-like growth factor (IGF) and Target of rapamycin (TOR) kinase signalling
to coordinate nutrition with cell growth, and steroid and neuropeptide hormones to
terminate feeding after a genetically encoded target weight is achieved. However, we still
understand little about how size is actually sensed, or how organ-intrinsic size controls
interface with whole-body physiology.

How growth, size and form are controlled during animal reproductive form. During metamorphosis, undifferenti-
development are problems that have entranced biologists ated progenitor cells, most of which are termed imaginal
for over a century. Recently, converging studies of cell cells, replace the older larval-specific cells, remodel the
growth and proliferation, pattern formation, endocrine existing organs and generate new ones, such as wings
regulation and evolution have generated new perspectives and eyes, de novo. Completely metamorphosing insects
to these problems. Progress in the insect model systems, (‘holometabolous’ insects) include common winged
as reviewed here, has been particularly noteworthy, as species such as flies (Diptera), butterflies and moths
genetic studies in Drosophila melanogaster have con- (Lepidoptera), ants and bees (Hymenoptera) and beetles
verged with classical endocrinological studies in other (Coleoptera). Insects do not grow as adults, and so their
insects to generate a hypothesis to explain body size, if final size can be considered, to a first approximation, as
not shape and form. The physiology of growth control a product of their growth rate during the larval phases
in insects is, of course, different to that in mammals, but and the duration of this growth period1–3.
the genes and signalling pathways that are involved are
surprisingly similar. For instance, insulin/insulin-like Critical weight and the ICG. One concept that is central
growth factor signalling (IIS) controls rates of cell growth, to this discussion is target size, which varies from spe-
nutrient use, cell size and body size in both flies and mice, cies to species and is therefore genetically determined.
and steroid hormone–nuclear receptor pairs regulate life- In holometabolous insects, the first manifestation of
stage transitions that affect growth in both. Therefore, target size is the commitment to metamorphosis. This
advances in insects, particularly in D. melanogaster, occurs sometime after the growing larva has passed
are influencing parallel studies in mice. Here I review a weight threshold that is termed the ‘minimum viable
recent findings that are relevant to the control of body weight’, operationally defined as the weight at which
size in D. melanogaster, with reference to other insects in larvae can develop into adults if food is completely
the many cases in which it is enlightening. withdrawn. This threshold was first recognized in
D. melanogaster by Beadle4 in the 1930s, and was distin-
Division of Basic Sciences, Insect development guished from a more biologically relevant one, termed
Fred Hutchinson Cancer Insects develop through a sequence of larval stages ‘critical weight’, some years later 5,6. Critical weight has
Research Center, 1100 called instars, which are interrupted by moults in which most often been defined as the weight after which feeding
Fairview Avenue North, the animal sheds its old exoskeleton and dons a new, no longer affects the time course to pupation, and has been
B2–152, Seattle, Washington
98109, USA.
larger one. Following a genetically specified number of studied most carefully in Manduca sexta7–9, a large moth
e-mail: bedgar@fhcrc.org moulting cycles, many insects enter a pupal stage, dur- that is a favourite system for studies of insect physiology. In
doi:10.1038/nrg1989 ing which they undergo metamorphosis to their adult, D. melanogaster, withdrawing food just as minimum

NATURE REVIEWS | GENETICS VOLUME 7 | DECEMBER 2006 | 907


© 2006 Nature Publishing Group
REVIEWS

Fat body viable weight is achieved delays the time course to pupa- Nutrition. Drosophila melanogaster larval development
A mesoderm-derived energy- tion5, and so mimimum viable weight occurs a few hours is complete after 4 days on rich food at 25oC, but can
storage organ that fulfils the earlier and is less than critical weight. Nevertheless, be extended to several weeks by restricting dietary pro-
functions that are assumed by because minimum viable weight is simpler to measure tein. Not surprisingly, larval growth can be arrested by
the liver and adipose tissues in
mammals.
experimentally, it has often been used interchangeably removing dietary protein completely. This treatment
with the term critical weight, especially in the Drosophila rapidly arrests cell growth and DNA replication in most
Adipokinetic hormones literature10. Critical weight is determined primarily by of the differentiated larval-specific tissues, but if the
Peptide hormones with genotype, and is clearly affected by numerous loci10,11, minimum viable weight has been attained, the progeni-
functions analogous to
as discussed below. In D. melanogaster, critical weight is tor cells that will form the adult continue to grow and
glucagons, produced by the
corpora cardiaca, a portion of
not significantly affected by diet4,6; slow-growing larvae, proliferate13, eventually generating a small but otherwise
the ring gland. These or larvae that are transiently starved, simply pass criti- normal fly. The fact that cells can grow within a starved
hormones stimulate the cal weight later. In M. sexta, however, poorer diets have animal indicates that the haemolymph (blood fluid)
mobilization of stored fat and been found to decrease the critical weight8. It is unclear in such animals is not critically depleted of nutrients.
carbohydrates from the fat
body upon starvation.
whether this difference is meaningful, or is due to dif- Indeed, D. melanogaster and other insects are known to
ferences in the way critical weight has been measured in maintain haemolymph nutrients when they are starved
Insulin-like peptides these two species. by mobilizing triglycerides and glycogen stored in the fat
(ILPs). Peptide hormones that It takes some time for an insect larva’s physiology to body. Nutrient mobilization is mediated by the induction
are homologous to vertebrate
sense critical weight and then initiate the behaviours that of metabolic neuropeptides called adipokinetic hormones
insulins and insulin-like growth
factors. ILPs are produced by
are associated with metamorphosis, including the cessa- (AKH), which are produced by the corpora cardiaca, a
medial neurosecretory cells in tion of feeding. Larvae continue to grow during this lag region of the neuroendocrine ring gland. AKHs function
the brain, as well as the gut period, referred to as the ‘interval to cessation of growth’ analogously to vertebrate glucagons14,15 and, together
and imaginal discs. These (ICG). Drosophila larvae grow so fast that they can more with insulin-like peptides (ILPs), are part of an endocrine
ligands bind the insulin
receptor and promote cellular
than quadruple their weight during the ICG if they are signalling system that allows the animal to coordinate
glucose import, energy storage cultured on rich food4. Therefore, although the normal rates of cell growth and changes in diet, with minimal
in the form of glycogen and dry weight for adult D. melanogaster in laboratory culture disruption of the developmental programme.
triglycerides, and cell growth. is about 290 µg, adults as small as 35 µg can be obtained
Drosophila melanogaster has
by starving the larvae to prevent growth during the ICG4. Insulin/insulin-like growth factor signalling. The insect
seven paralogous genes.
Orthologous genes in the silk
Growth trajectories during the ICG vary widely between IIS system (BOX 2) is highly homologous to that found
moth, Bombyx mori, are called species, and probably account for a significant amount in mammals. IIS activity promotes glucose import and
bombyxins. of species-specific size variation. M. sexta, for instance, nutrient storage by the fat body and other organs, fulfill-
barely doubles its mass during the ICG, even when food ing the homeostatic function of vertebrate insulins16–18.
Cell autonomous
If the activity of a gene has
is not limiting 8. In this Review, I first outline what is In this capacity, it affects feeding behaviour, lifespan and
effects only in the cells that known about the control of growth rates, as the rate of reproduction19. During development, IIS also regulates
express it, its function is said to growth during the ICG is an important parameter that cell growth, fulfilling the developmental function of
be cell autonomous; if it causes controls final body size. I then address how achieving the mammalian insulin-like growth factors (IGFs)20,21.
effects in cells other than (or in
critical weight leads to the onset of metamorphosis, IIS activity has been manipulated in various ways in
addition to) those that express
it, its function is cell non-
which is the essence of how growth-phase duration is D. melanogaster, using the Gal4–UAS system to overex-
autonomous. determined. Last, I delve into some unanswered ques- press genes in specific tissues, and the Flp–FRT system to
tions related to allometry, or how the proportions of the delete gene functions in specific tissues at defined times22.
various organs are controlled. These manipulations show that many IIS components
are not only essential for cell and organ growth, but are
Growth-rate control also sufficient to autonomously increase the growth rate
Growth rates during larval development are affected by of just about any cell type in D. melanogaster 23–26 (BOX 2).
nutrition, temperature (BOX 1), the density of animals in In whole animals, increased expression of several of
their environment12 and, of course, genotype. D. melanogaster’s seven ILPs can increase both larval
growth rates and adult size17,27,28, whereas ablation of the
small cluster of medial neurosecretory cells (mNSC) in
Box 1 | Temperature and body size the brain, which are the principal source of ILPs, reduces
growth rates and final body size16,17,28. Studies using a
Body size in ectotherms is generally affected by temperature80. In insects, lower temperature-sensitive allele of the insulin receptor (InR)29
temperatures decrease growth rates but actually increase final body size81. This affect
to regulate IIS at defined stages showed that the role of IIS
has been attributed to increased cell sizes rather than altered cell numbers82.
Davidowitz and Nijhout79 showed that, in Manduca sexta, growth rates increase linearly as a regulator of growth rate is general, but that its effects
with temperature, and proposed a simple explanation of why body and cell size on body size are limited to the ICG30. This supports earlier
nevertheless decrease. They show that the interval to cessation of growth (ICG) studies, in which changes in diet were used to show that
decreases markedly with increasing temperature, and that this more than counteracts the commitment to metamorphosis depends on achieving
the effect of faster growth rates. So, at lower temperatures, the longer ICG is presumed a critical size, rather than time or growth rate.
to allow more net growth, even though the growth rate is slower. This simple The Target of rapamycin (TOR) protein kinase is the
explanation fits the experimental data well79, but leaves open the question of why the best characterized, and arguably, the most important,
ICG is selectively affected at higher temperatures. This presumably derives from growth-regulatory target of insulin signalling23,31,32, at
the differential effects of temperature on feeding and growth, on the one hand, and the least in well-fed animals33,34. Studies in cultured cells show
kinetics of hormonal fluxes, on the other.
that, in addition to sensing nutritional state indirectly

908 | DECEMBER 2006 | VOLUME 7 www.nature.com/reviews/genetics


© 2006 Nature Publishing Group
REVIEWS

Box 2 | The insulin/insulin-like growth factor signalling system

ILPs Nutrition

Amino Glucose
acids
InR

PI3K AKT TSC1 ATP


a Starved
IRS TSC2 LKB1
PTEN AMPK
RHEB
b Fed Amino
acids
FOXO TOR

c Fed, ILP overexpression Transcription


S6K 4EBP TIF-IA
Growth Endocytosis
suppressors,
stress
response Translation Ribosomes Autophagy

The Drosophila melanogaster insulin/insulin-like growth factor signalling (IIS) system has been the subject of intense
genetic analysis, and is essentially similar to its human counterpart. It comprises a group of seven insulin-like peptides
(ILPs), a single insulin receptor (InR) gene, an insulin receptor substrate (IRS) protein called chico, the type IA phosphati-
dylinositol 3-kinase (PI3K), the lipid phosphatase PTEN and the protein kinases AKT/PKB and PDK1 (REF. 23). InR and its
downstream effectors are expressed ubiquitously, but the seven ILPs are expressed in specific tissues, presumably in
response to different inputs27,28. These ligands are nevertheless thought to mediate common cellular effects through the
same receptor. IIS affects growth by promoting the cellular import of glucose to enhance the cell’s energy supply, by
inhibiting the transcriptional activator FOXO33,83, which activates metabolic repressors such as 4EBP, and by maintaining
the activity of the Target of rapamycin (TOR), a conserved protein kinase. TOR is activated by the small GTPase RHEB
and promotes translational initiation, ribosome biogenesis, nutrient storage and endocytosis, and inhibits autophagy35,84.
In addition to regulation by IIS, TOR activity is also sensitive to cellular levels of amino acids and the ATP:ADP ratio,
which is an index of cellular energy levels. TOR senses energy levels through AMP-dependent kinase (AMPK), which
is activated by the LKB1 protein kinase, and inhibits the TSC1–TSC2 complex. It remains unclear how TOR senses
amino-acid levels. TIF-IA is a transcription factor that stimulates rRNA synthesis.
In the diagram, IIS factors are in blue, TOR-pathway components are shown in red and AMPK-pathway components are
shown in green.
The inset shows male flies, from the top down: part a, subjected to dietary restriction for protein; part b, raised on a rich
diet; and part c, subjected to constitutive, systemic ILP overexpression, also raised on rich food. Photographs courtesy of
S. Layalle, C. Géminard and P. Leopold, University of Nice.

through IIS, TOR responds cell autonomously to levels a crucial IIS transduction component38 (BOX 2). The first
of cellular amino acids, ATP and oxygen35. Although it is mechanism might dampen the cellular response to ILP
doubtful that cells inside an insect larva ever experience signalling when nutrient levels are high, whereas the
absolute amino-acid or glucose depletion as practised in second mechanism presumably blocks the potentially
cell-culture experiments, experiments in D. melanogaster harmful effects of inappropriate IIS when cells are criti-
indicate that certain larval cells that are resident in the cally starved for amino acids, glucose or oxygen. Another
fat body and perhaps other endocrine organs might use homeostatic mechanism involves the transcription fac-
TOR to sense the smaller fluctuations in haemolymph tor FOXO, which activates transcription of the InR, but
nutrients that accompany changes in diet36,37. is also repressed by IIS activity through AKT39.
There is a finely tuned, elaborate feedback network Numerous observations underscore the importance
Intracellular second between IIS, TOR and their nutritional inputs. Glucose of IIS and TOR signalling as a nutrient-response sys-
messenger import, which is controlled by IIS, probably affects cel- tem. Starvation represses the expression of several of
A signalling molecule in cells, lular ATP pools, which control TOR activity. TOR also the ILPs27,28,36, depletes the IIS second messenger PIP3
the concentration of which feeds back on IIS in at least two ways. First, TOR nega- (REF. 24), and inhibits the activity of the TOR target S6K
changes on the binding of an
extracellular ligand to a
tively regulates ISS through feedback between the kinase (REF. 40). Consistent with this, a silk moth ILP, bombyxin,
plasma-membrane-bound S6K and insulin receptor substrate (IRS) proteins37, is depleted from the haemolymph on starvation41.
receptor. and, second, TOR activity is required to activate AKT, Perhaps most significant of all, genetically activating IIS

NATURE REVIEWS | GENETICS VOLUME 7 | DECEMBER 2006 | 909


© 2006 Nature Publishing Group
REVIEWS

Box 3 | Another function for juvenile hormone into starved Bombyx mori (silk moth) larvae is sufficient
to stimulate the secretion of bombyxin 41. Therefore,
Juvenile hormone (JH) has been well characterized as a suppressor of although it is clear that the IIS system somehow senses
prothoracicotropic hormone (PTTH) and 20-hydroxyecdysone (20E) release, but in
dietary protein, the underlying mechanism remains
Manduca sexta it also interacts with the insulin/insulin-like growth factor (IIS) system
obscure. It has been suggested that ILP secretion from
to suppress the formation and growth of late-forming imaginal discs45. Such discs,
unlike the extensively studied early-forming wing and eye discs of Drosophila the mNSCs might be regulated by dietary protein, but it
melanogaster, undergo most of their growth after the onset of metamorphosis and seems equally likely that some other organ senses amino
pupariation, using nutrients stored in earlier stages. Drosophila melanogaster has acids and promotes IIS activity indirectly, through
analogous sets of imaginal cells: examples include the nests of histoblasts that secondary factors.
generate the adult abdominal epidermis and the islands of imaginal cells that reside
in the larval gut. The formation and growth of the late-forming eye and leg discs in The fat body. This other nutrient-sensing organ could
M. sexta are arrested by starvation before attainment of critical weight, but they will be the fat body. Although the fat body has not been
form and grow in starved animals from which the corpora allata, the JH-producing reported to produce ILPs27,28, organ-culture experi-
organ, has been excised45. These effects are ecdysone independent, suggesting that
ments indicate that it does produce some sort of growth
JH might function directly as a negative growth factor, at least on eye, leg and
factor13,46, and that this might be nutrient dependent.
wing imaginal cells. However, the in vivo application of a stable JH mimetic does
not suppress the formation or growth of the late discs85, indicating that the growth- Recent reports show that specifically suppressing
suppressive action of JH is eventually overcome by nutrient-dependent signals (for metabolism in the fat body by inhibiting amino-acid
example, IIS) in feeding animals. The molecular targets by which JH suppresses import, genetically squelching TOR activity or activat-
imaginal cell growth remain unknown, as does the general mode of JH signal ing FOXO is sufficient to non-autonomously suppress
transduction86. Whether JH modulates growth in tissues other than the late-forming IIS activity in both larvae36 and adults47. The nutrition-
discs or in insects other than M. sexta are also open questions. dependent production of an IIS cofactor by the fat body
is one attractive explanation for these results (FIG. 1).
The role of TOR in the fat body might be either to sense
or TOR signalling can bypass the known cellular effects amino-acid and ATP pools, or to promote the synthe-
of starvation, including the arrest of cell growth and sis of this putative IIS cofactor, or both. A putative
DNA replication in the larval-specific tissues24,42, and the insulin-binding factor, acid labile subunit (ALS), which
induction of autophagy in the fat body 43,44. These obser- is produced by the fat body only in feeding animals36, is a
vations indicate that IIS and TOR activity are suppressed candidate for this missing growth factor. Other can-
in the larval-specific tissues by starvation, and that this didates are the chitinase-related imaginal disc growth
systemically mediates a set of cellular responses that factors (IDGFs)48 and adenosine-deaminase-related
allow insect larvae to cope with fluctuating nutritional growth factor D (ADGFD)49,50, which are mitogens that
conditions. are expressed by the fat body.
As noted above, adult progenitor cells in the brain
and imaginal discs are relatively insensitive to changes Ecdysone and the prothoracic gland. Virtually all bio-
in diet. In D. melanogaster, growth and division of these logical regulatory systems have elements of negative
cells is not arrested in animals that are starved after feedback that keep them from running wild. An impor-
reaching critical weight. These cells might require less tant mode of negative feedback in body-size control
of the ILP stimulus to maintain high rates of anabolic seems to involve 20-hydroxyecdysone (20E)51–53, the ster-
metabolism, or they might produce their own ILPs. oid hormone that controls both moulting and the onset
Indeed, starvation-insensitive ILP expression has of metamorphosis in all insects54,55. Recent studies found
been documented in both locations27,28. The nutrition- that 20E activity is positively controlled by IIS activity
independent growth of progenitor cells in D. melanogaster in the prothoracic gland, a sector of the ring gland that
larvae probably reflects the animal’s life strategy, which is produces the steroid hormone ecdysone, which is the
to prioritize nutrient utilization for generating reproduc- 20E precursor51–53. Suppressing IIS in the prothoracic
tively capable adults. Of course, this strategy succeeds gland reduces 20E activity and increases adult body size,
only if an animal has achieved minimum viable weight whereas increasing IIS has the opposite effect. How IIS
before it is starved. Animals that are starved before this stimulates ecdysone production is still unclear, but the
do not have sufficient stored nutrients to support growth same relationship has been documented in other organs
of the imaginal cells to term, and they eventually perish. in mosquitos and silk moths19, and therefore is probably
In insects other than D. melanogaster, the developmen- general. Other factors that affect cell growth, such as Myc
tal strategy for achieving the same life strategy can be and cyclin D–CDK4, do not seem to stimulate ecdysone
different, and this is reflected in some of the physi- production by the prothoracic gland, whereas activating
20-hydroxyecdysone ological differences that have been noted. Starvation of the Ras–Raf pathway does53. The explanation for this is
(20E). The insect moulting
M. sexta larvae, for instance, rapidly arrests growth of the not clear, but one possibility is that both IIS and Ras–Raf
hormone, a steroid. The
pro-hormone ecdysone is imaginal discs45 (BOX 3). signalling promote specific aspects of metabolism that
produced by the prothoracic Because dietary protein, and not sugar, is required are required for steroid synthesis.
gland. Ecdysone is then for cell growth, one might expect the mNSCs, which It was expected that the body-size effects that result
converted into active 20E produce the ILPs, to be sensors of dietary protein. from altering the prothoracic gland would be due to
by the fat body, malpighian
tubules (analogous to the
However, experiments indicate that it is sugar, not changes in the timing of the onset of metamorphosis,
kidneys) and possibly protein, that mNSCs require to express ILP mRNA in which is triggered when the prothoracic gland releases a
the epidermis. D. melanogaster 27,36. Consistently, injection of glucose large pulse of ecdysone (BOX 4). Indeed, extreme changes

910 | DECEMBER 2006 | VOLUME 7 www.nature.com/reviews/genetics


© 2006 Nature Publishing Group
REVIEWS

although high 20E levels inhibit cell proliferation, low


JH PTTH levels of 20E or ecdysone are required for the growth
ILP and proliferation of some larval cells in culture56,57. The
ability of ecdysone to antagonize IIS is therefore either
JHE CA
level dependent, or perhaps cell-type specific. Further
PG mNSCs studies addressing the interface of ecdysone, IIS and
TOR signalling are required to clarify this interesting
and unexpected connection.
Critical weight 20E ? ILP Nutrition
Cell-size control
Changes in cell size account for a significant amount of
Fat body the variation in body size that is seen in D. melanogaster,
both in wild populations58 and experimental situa-
tions32,33. The size of proliferating cells is determined
?
by their relative rates of growth and division. I have
Cessation of discussed growth-rate control, but how are division
feeding Peripheral rates regulated? Rates of cell division in the imaginal
metamorphosis tissue growth
discs are clearly growth dependent: when growth rates
Figure 1 | Endocrine communication between the organs involved in growth are depressed by lack of nutrition, IIS or TOR signalling,
control. Organs and cells are denoted as ovals, hormones are denoted as coloured cell division slows down too. However, experimental
boxes and effects are denoted in grey boxes. Nutrition allows growth of tissues manipulations that upregulate cell growth rates generally
throughout the body, and also stimulates the production of insulin-like peptides do not increase the speed or number of cell divisions
(ILPs) by medial neurosecretory cells (mNSCs) in the brain. ILPs activate the insulin/ in these organs23,59, at least when the experiments are
insulin like growth factor–Target of rapamycin (IIS–TOR) system and promote performed on rich food. Therefore, there seems to be
growth systemically. The fat body also senses nutrition, probably using TOR, and an intrinsic developmental programme that determines
produces poorly characterized growth factors that might potentiate IIS. IIS to the the maximal rate of division of the imaginal cells. As
prothoracic gland (PG) promotes the production of 20-hydroxyecdysone (20E) by
discussed below, the final number of cells in each organ
this organ, providing negative feedback that supresses cell growth and, eventually,
the cessation of feeding and the onset of metamorphosis. Juvenile hormone (JH), is also developmentally programmed, largely inde-
juvenile hormone esterase (JHE) and prothoracicotropic hormone (PTTH) provide pendently of growth rate. It is partly because of these
another negative-feedback loop. JH is produced by the corpora allata (CA) in young independent controls on cell division rates and cell
growing animals, and suppresses 20E levels by inhibiting PTTH release. However, numbers that nutrition, IIS and TOR signalling have
when the growing animal achieves critical weight, rising levels of JHE (produced such limited effects on final cell numbers in the adult
by the fat body and other organs) degrade JH, leading to the release of PTTH, but such profound effects on cell size: if cell growth rates
which triggers ecdysone production. This in turn suppresses growth, ends feeding are increased without coincidentally increasing division,
and causes metamorphosis. cell size increases.
Adult insects consist almost completely of non-
proliferating cells, and so the sizes of their cells are
in IIS activity in the prothoracic gland do change the determined simply by the relative rates of synthesis,
timing of metamorphosis, and therefore probably also storage and turnover of macromolecules, metabolites
alter the duration of the ICG52,53. Surprisingly, how- and water. Accordingly, most of the genes that affect cell
ever, Colombani51 and Mirth52 found that more subtle size in adults alter metabolic balance in some way. This
changes in IIS in the prothoracic gland clearly affected includes the IIS- and TOR-signalling genes that were
rates of larval growth from early development, with- discussed above (BOX 2), and a few others such as Myc
out affecting developmental timing. Closer inspection and the cyclin D–CDK4 kinase59–61, which affect post-
revealed that 20E and its receptor, EcR, antagonize the mitotic cell growth. Like TOR, Myc is a potent stimulator
ability of ILP signalling to activate PI3K and AKT, and of ribosome biogenesis and protein synthesis61, but
to suppress the nuclear localization of FOXO51 (BOX 2). seems to be regulated by IIS- and TOR-independent,
The effect on FOXO was found to be cell autonomous, non-nutrition-related developmental signals. Cyclin D–
at least in the fat body, implying that IIS and 20E proba- CDK4 is a non-essential regulator of cell growth that has
bly have antagonistic effects on growth and metabolism a separate function in controlling cell-cycle progression
at the cellular level. Consistent with the role of the fat through the conserved E2F–RB pathway 62. Cyclin D–
body in whole-body growth control, ecdysone signal- CDK4 upregulates mitochondrial activity by unknown
ling to the fat body has non-cell-autonomous, systemic mechanisms, and requires intact mitochondrial function
effects on growth. For instance, specifically suppressing to stimulate cell growth60. Genetic tests show that each of
ecdysone signalling to the fat body increases overall these gene functions affects not only cell size, but body
growth rates and final body size, without affecting size as well, and therefore they are potential targets for
development time. An attractive explanation for this evolutionary selection of body size.
is that ecdysone signalling suppresses the production Population-genetics studies have also documented
of nutrient-dependent growth factors by the fat body significant differences in cell size between large and
(FIG. 1). However, the role of 20E as a growth factor is far small strains of D. melanogaster 58,63, although the
from simple; observations in other insects show that, genetic basis of these changes has only been guessed

NATURE REVIEWS | GENETICS VOLUME 7 | DECEMBER 2006 | 911


© 2006 Nature Publishing Group
REVIEWS

at (BOX 5). It is logical to assume that adult cell sizes, Growth-phase duration
and even organ sizes, might be determined in part The timing of the cessation of feeding and the onset of
by differentially programmed expression of IIS- and metamorphosis is a critical determinant of body size
TOR-pathway components58,64, and at least one recent in insects. This transition is controlled, at least in part,
study supports this view with data18. However, virtu- by a physiological network of hormones that circulate
ally any factor that alters the balance between anabolic in the animal’s haemolymph54 (FIG. 1). As outlined in
and catabolic metabolism could be expected to act as a BOX 4, the growth phase is terminated by a hormonal cas-
determinant of cell size. cade, which starts with the attainment of critical weight
and ends with the secretion of high levels of ecdysone,
triggering the cessation of feeding and pupariation. In
M. sexta, the signal that critical weight has been reached
Box 4 | Hormonal controls that time metamorphosis
is relayed by a drop in juvenile hormone (JH), followed
by secretion of prothoracicotropic hormone (PTTH),
Weight which triggers ecdysone production (FIG. 1). The same
JH is assumed to be true in D. melanogaster, although
20E
PTTH comprehensive data are lacking.

Models for size control. Davidowitz et al.8, Nijhout et al.9


and Shingleton et al.18 have elaborated on the early find-
ings of Beadle4, Bakker5 and others to develop models for
the physiological control of body size. These models
describe in specific terms the interplay between growth
rate, critical weight and the control of JH, PTTH and
ecdysone during the final larval instar of D. melanogaster
or M. sexta development. Both have three critical param-
eters: the growth rate, the critical weight and the length
ICG
of the ICG (FIG. 2). After critical weight is attained, the
1st moult 2nd moult Critical Cessation Pupariation animal continues to grow until a spike of 20E causes
weight of growth the cessation of feeding and the onset of pupariation
attained
and metamorphosis (BOX 4). Because larval growth is
Classic work in several insects showed that the cessation of feeding and onset of essentially exponential, growth during the ICG can be
metamorphosis are triggered by pulses of circulating 20-hydroxyecdysone (20E), the substantial, accounting for about half of the peak larval
same steroid hormone that triggers earlier moults. The graph is an idealized mass in M. sexta and most of it in D. melanogaster, pro-
schematic that represents hormone fluxes in the haemolymph of a growing larva. vided that growth occurs on rich food. If growth is slowed
Attainment of critical weight suppresses juvenile hormone (JH) production, leading or stopped during the ICG by poor diet, loss of insulin sig-
to prothoracicotropic hormone (PTTH) production, ecdysone secretion and the
nalling or crowding, for instance, then the final weight of
cessation of feeding. A large fraction of total growth occurs in the interval
to cessation of growth (ICG).
the animal will be much closer to its critical weight.
Insects produce ecdysone in their prothoracic glands, which sense multiple inputs. This model has been carefully tested both experimen-
Insulin/insulin-like growth factor (IIS) and Ras–Raf signalling to the prothoracic gland tally 9,18 and by computer-based simulations9, and seems
affect 20E levels51–53, suggesting that 20E production is controlled by nutritional and to be a useful framework for understanding the effects
cell-specification signals, respectively. In Bombyx mori, 20E production can also be of nutrition, IIS and ecdysone signalling on body-size
controlled by direct innervation of the prothoracic gland87, presumably from the brain. control in Diptera and Lepidoptera. It provides a simple
However, the pulses of 20E that trigger moulting, the cessation of feeding and explanation for the effects of diet and IIS , because these
pupariation are triggered by PTTH, which is secreted by cells in the nearby brain. In some factors have big effects on growth rates during the ICG
insects, such as milkweed bugs (Oncopeltus fasciatus) and blood-sucking Hemitpera but relatively insignificant effects on critical weight4,8,18.
(Rodnius prolixus and Dipetalogaster maximus), PTTH secretion is triggered by a size-
However, the current paradigm has limitations. One is
measuring mechanism that involves stretch receptors in the abdomen. When the
abdomen is distended by feeding, these stretch receptors signal to the brain to release
that it applies only to the last instar of larval develop-
the hormone2. However, this mechanism is not universal. In Onthophagus taurus, a ment, and leaves open the important question of how
beetle, PTTH secretion is triggered when the larva’s food supply is exhausted88, and in critical weight is determined and sensed, and how this
Manduca sexta, a moth, PTTH secretion is inhibited by JH, a sesquiterpene hormone that information is then relayed to the ring gland to end JH
is present through early development and then cleared from the haemolymph when the production and induce PTTH and ecdysone production.
larva reaches a critical weight86. The loss of JH is due, in part, to a rise in the activity of This question would seem to be at the heart of overall
juvenile hormone esterase (JHE), an enzyme that breaks down JH and is produced by the size control, at least in Diptera and Lepidoptera.
fat body in feeding animals72. In M. sexta, JH levels fail to drop if larvae are starved before
attaining critical weight, and experimental application of JH can block pupation and Sensing critical weight. So far, little is known about what
extend the period of larval growth, giving rise to oversize larvae. In Drosophila
parameter is actually ‘read’ as critical weight. However,
melanogaster, excess JH also prolongs the final (third) instar, but this treatment does
not lead to larger animals89. PTTH secretion in M. sexta, is further regulated by a
there are some thought-provoking observations to con-
photoperiodic ‘gate’90, such that it can only be released during an 8-hour window each sider. For instance, early studies in several insects showed
day. Therefore, the time of day at which the larva clears its JH is another variable that that the presence of growing or regenerating discs delays
affects body size (FIG. 2). This control has not been demonstrated in D. melanogaster, pupariation65–67, indicating that growing imaginal tissues
although recent work hints at its existence52. Figure after REFS 54,91. generate a signal that maintains JH at high levels until

912 | DECEMBER 2006 | VOLUME 7 www.nature.com/reviews/genetics


© 2006 Nature Publishing Group
REVIEWS

Box 5 | Evolution of Drosophila melanogaster body size suppresses cell growth51–53, rising levels would be expected
to first slow growth in peripheral tissues, and then trigger
Latitudinal clines in body size have been documented between Africa and Europe, from the cessation of feeding. A second analogous feedback
the Baltic to Central Asia, along the east coast of North America, and in Australia and
mechanism can be envisioned for the clearing of JH.
Southern Africa. Flies with larger adult body size, faster development rate and higher
In this case, exponential growth of the fat body, and a
fecundity reside at higher latitudes, corresponding to climates with cooler temperatures
and harsher winters58. Conversely, smaller body size, slower development and decreased concurrent exponential rise in the capacity of this organ
fecundity, but increased starvation resistance and longevity, are found in more to produce juvenile hormone esterase (JHE)72, might
equatorial populations. These adaptations are thought to reflect the different selective help to clear JH and trigger the cascade that leads from
pressures in tropical versus temperate regions; fast development and thermotolerance PTTH to the production of 20E at a critical size. Similar
are essential when flies must overwinter and then develop and breed quickly in the feedback loops can be envisioned in which fat-body
spring, whereas longevity and the ability to survive crowding and starvation are growth enhances 20E signalling through its ability to con-
advantageous in warm climes. The main determinant of body-size variation in natural vert the ecdysone prohormone to active 20E, or through
populations seems to be altered cell numbers (at least in the wings where cell numbers the production of growth factors that augment IIS activ-
are most often tallied), but changes in cell size are also common63,92. Increased metabolic
ity in the prothoracic gland36,48,50. Consistent with these
rate and efficiency have also been correlated with larger body size58.
ideas, increased IIS activity in the fat body causes not only
The evolution of larger body size has been reproduced in several laboratories93,94,
and most of the parameters that are seen to change in the wild, such as cell size, increased fat-body growth, but precocious cessation of
development rate and critical weight were found to be affected in a similar way. feeding24. The combined function of these sorts of negative-
When small and large varieties of D. melanogaster are reared in identical conditions in feedback loops might explain, in part, why larvae do not
the laboratory, their size differences and differences in critical weight breed true10,94, grow past critical size. However, these mechanisms have
and are therefore genetically encoded. Several attempts have been made to not been directly tested in the laboratory. How much
genetically map QTLs that affect body size58,95, but these have so far failed to define of the haemolymph ILP is contributed by the discs and
the specific genes involved. Various authors have suggested that components of the other tissues is unknown, studies of the fat body as a
insulin/insulin like growth factor–Target of rapamycin (IIS–TOR) signalling system signalling centre have just begun, and the few studies
could be important targets for selection of body size58,64, and most of the available
of pupariation timing that have been published73 do not
data are consistent with this. Other factors, such as those that determine sensing of
directly address these ideas. A comprehensive genetic
critical weight, the kinetics of juvenile hormone (JH) decline and prothoracicotropic
hormone (PTTH) release, and the myriad determinants of the interval to cessation of analysis of the mechanisms that sense critical weight and
growth (ICG) are also probable targets of selection. control pupariation timing in D. melanogaster should add
significantly to the already rich physiological literature
from M. sexta, B. mori and other large insects.
their growth reaches some milestone, which could be a Another limitation of the current paradigm (FIG. 2)
metric for critical weight. Consistent with this notion, is its assumption that adult size is proportional to peak
mutations that disrupt cytoarchitecture in the imaginal larval size. Although this condition is consistent with
disc cells, thereby allowing them to grow indefinitely, much of the data from D. melanogaster and M. sexta, the
suppress 20E levels, prolong larval growth and, in some available data also indicate that the maximal sizes of
cases, generate oversize pupae65,67,68. But, although the the adult organs are ultimately controlled by organ-
idea that signals that originate from the discs act on the intrinsic signalling, rather than as a function of total body
endocrine signalling centres to time metamorphosis size or stored nutrients (see below). If this is true, there is
has existed for many years, such signals have never been probably an upper limit on adult body size that is geneti-
identified. Moreover, several facets of larval biology are cally encoded, and that cannot be exceeded by simply
hard to reconcile with this idea. For instance, the vari- increasing growth and nutrient storage during the ICG.
ous imaginal tissues grow and mature asynchronously, It will be interesting to see whether increasing nutrient
and many continue to grow long after critical weight storage by larvae to three or four times the normal level
is achieved and even after the cessation of feeding. In results in proportional increases in adult size, or whether
insects other than flies, post-pupation disc growth is, in nutrients that are stored in excess of what is required to
fact, the norm (BOX 3). These observations indicate that achieve a maximum, genetically programmed target size
disc growth per se does not inhibit metamorphosis. A are simply discarded. Given the current state of the field,
Juvenile hormone second confounding observation is that mutant larvae this should soon be possible in D. melanogaster.
(JH). A sesquiterpene produced that lack imaginal discs altogether can pupariate69,70,
by the corpora allata, which is a although this occurs later and at a smaller size than nor- Cell numbers and allometry
portion of the ring gland. JH
promotes juvenile (larval)
mal71. Therefore, although disc growth might provide a So far, experimental manipulations of nutrition and
development and inhibits ‘growth checkpoint’ in some insects, there must be an hormonal signalling have succeeded in reducing the
metamorphosis. It is degraded, underlying disc-independent mechanism that can sense weight of D. melanogaster adults to about 15% (REF. 4), or
in part, by juvenile hormone critical weight and trigger pupariation. increasing it to about 150% (REF. 28), of what is normally
esterase, which is produced by
Despite these puzzling discrepancies, recent find- attained in the laboratory when the flies are fed on rich
the fat body and malpighian
tubules (kidney). ings36,51–53 support at least one potential size-sensing food. Much of this variation is due to changes in cell
mechanism that is consistent with most of the literature size, rather than cell numbers, which in D. melanogaster
Prothoracicotropic hormone and anecdotal accounts. If the imaginal discs, gut and have been decreased by genetic manipulation to about
(PTTH). A neuropetide that is salivary glands produce ILPs27, the exponential growth 80% (REFS 26,33), or increased to about 120% of nor-
secreted by cells in the brain.
PTTH stimulates ecdysone
of these organs might cause an exponential increase mal28,51. However, natural variation in cell numbers
production by the prothoracic in IIS to the prothoracic gland, and thereby promote between species of Drosophila, or even subpopulations of
gland. its increasing production of ecdysone. Because 20E D. melanogaster, are greater than this58,63,74. If one considers

NATURE REVIEWS | GENETICS VOLUME 7 | DECEMBER 2006 | 913


© 2006 Nature Publishing Group
REVIEWS

PCG ICG

Passed Yes Yes Yes Secrete PTTH Stop feeding


Last larval JH titres JH cleared from Photoperiodic
critical Stop growing
instar decline haemolymph? gate open? and ecdysone
weight? Begin metamorphosis

No No No
Continue Continue Continue
growing growing growing

Figure 2 | Control of growth during the final larval instar. The figure shows the checkpoint that governs the interval
to cessation of growth (ICG), and the two checkpoints within the ICG before the cessation of growth. JH, juvenile hormone;
PCG, growth prior to critical weight; PTTH, prothoracicotropic hormone. Figure modified from REF. 79. See also REF. 9.

more divergent species such as D. melanogaster and affect target size. Such short-range signals are commonly
M. sexta, or mice and elephants, differences in cell thought of as regulators of allometry, but considering how
numbers are vastly greater than differences in cell size. much communication there is between the various organs
Therefore, from an evolutionary standpoint, the control (FIG. 1), it would be surprising if short-range signals did
of cell numbers is a more important determinant of body not feed back at multiple levels on the endocrine signals
size than the control of growth rates or cell size. The that affect critical-weight sensing, growth rates and the
control of cell numbers remains something of a mystery, duration of the growth period. Studies in beetles that
and one that will probably not be resolved by studies document growth-regulatory interactions between differ-
of nutrition, IIS/TOR signalling and endocrinology. ent imaginal discs are consistent with such feedback64,78,
There are some clues, however. and the discovery that Ras–MAPK signalling in the
One telling observation is that when imaginal discs prothoracic gland affects 20E levels, pupariation timing
are cultured for long periods in growth-permissive and body size53 provides a tangible molecular example.
environments, such as young larvae or the abdomens Therefore, in considering the evolution of body size, there
of adults, they grow only to their normal target size and are few things that can be ignored.
then stop75. Therefore, target cell number in these adult
organs is essentially an organ-autonomous property. The Future directions
same can be assumed of the CNS, which has a virtually The work discussed here is an inspiring example of
invariant pattern of cell divisions13,76. The prevailing how the convergence of two previously separate fields
explanation for this is that organ size (and therefore, of study — Dipteran genetics and Lepidopteran physi-
final cell numbers in the adult) are determined primarily ology — can yield a new paradigm for understanding
by the way short-range, organ-autonomous (paracrine) a sizeable biological issue. However, the synthesis is
signalling systems, such as the Wnt, bone morphogenic incomplete, because although most of the crucial organs
protein (BMP), hedgehog (HH), epidermal growth and gene systems that control body size might have been
factor (EGF) and Notch pathways, are deployed dur- identified, the way these units communicate (FIG. 1) is still
ing development. The specifics of local cell signalling, poorly understood. Moreover, the size-control paradigm
combined with organ-specific codes of transcriptional as it stands (FIG. 2) refers to a rather narrow window of
regulators (for example, Hox genes) seem to give each time — the last instar of larval development — and aims
organ a genetically predetermined target size, which can to explain only the three- to fourfold variations in size
be modulated over only a modest range by nutritional that have been achieved experimentally, and that are seen
and endocrine inputs. Many genetic experiments with within species in the wild. Understanding the much larger
short-range signals and selector-type transcription fac- variations in body size between species that occur dur-
tors support this view (see REFS 1,3,59 for reviews). To ing evolution will require new concepts and approaches.
take an example from mammals, the evolution of wings Future studies focused on the functional connections
in bats is thought to derive from increased proliferation between intrinsic size control in the imaginal tissues and
of the chondrocytes that lay down the finger bones. This critical-weight sensing by the whole animal are clearly
has made the bat’s fingers, which form the wing, many important. It would be best if these studies were carried
times longer than the fingers of their mouse-like ances- out using combined genetic and physiological tools in
tors. Increased chondrocyte proliferation in bat forelimb the same organism, to avoid the pitfalls of comparing
digits seems to derive from prolonged expression of a systems (for example, D. melanogaster, M. sexta and
BMP, which functions as a local growth factor77. B. mori) that have significant physiological differences.
In flies, short-range BMP signalling controls the tar- Incorporating evolutionary genetics is also an important
get size and shape of most of the organs discussed here, goal, but this will be challenging, because comparative
including imaginal discs, the ring gland, the fat body and genomics without experimentation is unlikely to explain
various regions of the brain. Other local signals such as the intricacies of anatomy and physiology that, to a large
wingless (WG, a Wnt), HH, Notch, and the various ligands extent, control how an organism regulates its growth rate
and receptors that feed into the Ras–MAPK pathway also and determines its final size.

914 | DECEMBER 2006 | VOLUME 7 www.nature.com/reviews/genetics


© 2006 Nature Publishing Group
REVIEWS

1. Stern, D. L. & Emlen, D. J. The developmental 23. Oldham, S. & Hafen, E. Insulin/IGF and target of 46. Davis, K. T. & Shearn, A. In vitro growth of imaginal
basis for allometry in insects. Development 126, rapamycin signaling: a TOR de force in growth control. disks from Drosophila melanogaster. Science 196,
1091–1101 (1999). Trends Cell Biol. 13, 79–85 (2003). 438–440 (1977).
A creative and authoratative review that 24. Britton, J. S., Lockwood, W. K., Li, L., Cohen, S. M. & 47. Hwangbo, D. S., Gershman, B., Tu, M. P., Palmer, M. &
summarizes the most important findings that are Edgar, B. A. Drosophila’s insulin/PI3-kinase pathway Tatar, M. Drosophila dFOXO controls lifespan and
relevant to allometry in insects. coordinates cellular metabolism with nutritional regulates insulin signalling in brain and fat body.
2. Nijhout, H. F. The control of body size in insects. conditions. Dev. Cell 2, 239–249 (2002). Nature 429, 562–566 (2004).
Dev. Biol. 261, 1–9 (2003). 25. Oldham, S. et al. The Drosophila insulin/IGF receptor 48. Kawamura, K., Shibata, T., Saget, O., Peel, D. &
3. Day, S. J. & Lawrence, P. A. Measuring dimensions: controls growth and size by modulating PtdInsP(3) Bryant, P. J. A new family of growth factors produced
the regulation of size and shape. Development 127, levels. Development 129, 4103–4109 (2002). by the fat body and active on Drosophila imaginal disc
2977–2987 (2000). 26. Bohni, R. et al. Autonomous control of cell and organ cells. Development 126, 211–219 (1999).
4. Beadle, G., Tatum, E. & Clancy, C. Food level in relation size by CHICO, a Drosophila homolog of vertebrate 49. Zurovec, M., Dolezal, T., Gazi, M., Pavlova, E. &
to rate of development and eye pigmentation in IRS1–4. Cell 97, 865–875 (1999). Bryant, P. J. Adenosine deaminase-related growth
Drosophila melanogaster. Biol. Bull. 75, 447–462 27. Brogiolo, W. et al. An evolutionarily conserved factors stimulate cell proliferation in Drosophila by
(1938). function of the Drosophila insulin receptor and depleting extracellular adenosine. Proc. Natl Acad.
The classic paper on critical weight. insulin-like peptides in growth control. Curr. Biol. 11, Sci. USA 99, 4403–4408 (2002).
5. Bakker, K. Feeding period, growth, and pupation in 213–221 (2001). 50. Dolezal, T., Dolezelova, E., Zurovec, M. & Bryant, P. J.
larvae of Drosophila melanogaster. Entomol. Epis. 28. Ikeya, T., Galic, M., Belawat, P., Nairz, K. & Hafen, E. A role for adenosine deaminase in Drosophila larval
Appl. 2, 171–186 (1959). Nutrient-dependent expression of insulin-like peptides development. PLoS Biol. 3, e201 (2005).
6. Robertson, F. W. The ecological genetics of growth in from neuroendocrine cells in the CNS contributes to 51. Colombani, J. et al. Antagonistic actions of ecdysone
Drosophila. 6. The genetic correlation between the growth regulation in Drosophila. Curr. Biol. 12, 1293 and insulins determine final size in Drosophila. Science
duration of the larval period and body size in relation (2002). 310, 667–670 (2005).
to larval diet. Genet. Res. 4, 74–92 (1963). The most comprehensive of several studies that 52. Mirth, C., Truman, J. W. & Riddiford, L. M. The role of
7. Nijhout, H. F. & Williams, C. M. Control of moulting show that increased ILP expression can increase the prothoracic gland in determining critical weight
and metamorphosis in the tobacco hornworm, adult body size in D. melanogaster. for metamorphosis in Drosophila melanogaster.
Manduca sexta (L.): growth of the last-instar larva and 29. Chen, C., Jack, J. & Garofalo, R. S. The Drosophila Curr. Biol. 15, 1796–1807 (2005).
the decision to pupate. J. Exp. Biol. 61, 481–491 insulin receptor is required for normal growth. 53. Caldwell, P. E., Walkiewicz, M. & Stern, M. Ras activity
(1974). Endocrinology 137, 846–856 (1996). in the Drosophila prothoracic gland regulates body
8. Davidowitz, G., D’Amico, L. J. & Nijhout, H. F. 30. Shingleton, A. W. Body-size regulation: combining size and developmental rate via ecdysone release.
Critical weight in the development of insect body genetics and physiology. Curr. Biol. 15, R825–R827 Curr. Biol. 15, 1785–1795 (2005).
size. Evol. Dev. 5, 188–197 (2003). (2005). References 51–53 demonstrate that IIS activity
An important study that defines critical weight in 31. Zhang, H., Stallock, J. P., Ng, J. C., Reinhard, C. & in the prothoracic gland stimulates ecdysone
M. sexta, and shows that it is affected by diet. Neufeld, T. P. Regulation of cellular growth by the production and thereby reduces adult size.
9. Nijhout, H. F., Davidowitz, G., Roff, D. A. Drosophila target of rapamycin dTOR. Genes & Dev. Reference 53 also shows that Ras–MAPK activity
A quantitative analysis of the mechanism that 14, 2712–2724 (2000). has a similar effect.
controls body size in Manduca sexta. 5, e16 (2006). 32. Oldham, S., Montagne, J., Radimerski, T., Thomas, G. 54. Riddiford, L. M. in The Development of Drosophila
A quantitative and theoretical treatise that & Hafen, E. Genetic and biochemical characterization melanogaster (eds Bate, M. & Martinez Arias, A.)
rigorously evaluates the model that is shown in of dTOR, the Drosophila homolog of the target of 899–939 (Cold Spring Harbor Lab. Press, Cold Spring
Figure 2. rapamycin. Genes Dev. 14, 2689–2694 (2000). Harbor, 1993).
10. De Moed, G. H., Kruitwagen, C. L. J. J., De Jong, G. & 33. Junger, M. A. et al. The Drosophila forkhead 55. Riddiford, L. M., Cherbas, P. & Truman, J. W.
Scharloo, W. Critical weight for the induction of transcription factor FOXO mediates the reduction in Ecdysone receptors and their biological actions.
pupariation in Drosophila melanogaster: genetic and cell number associated with reduced insulin signaling. Vitam. Horm. 60, 1–73 (2000).
environmental variation. J. Evol. Biol. 12, 852–858 J. Biol. 2, 20 (2003). 56. Nijhout, H. F. & Grunert, L. W. Bombyxin is a growth
(1999). 34. Kramer, J. M., Davidge, J. T., Lockyer, J. M. & Staveley, factor for wing imaginal disks in Lepidoptera. Proc.
This paper discusses how critical weight is B. E. Expression of Drosophila FOXO regulates growth Natl Acad. Sci. USA 99, 15446–15450 (2002).
measured in D. melanogaster, and shows that and can phenocopy starvation. BMC Dev. Biol. 3, 5 57. Champlin, D. T. & Truman, J. W. Ecdysteroid
it is affected by genotype. (2003). coordinates optic lobe neurogenesis via a nitric oxide
11. D’Amico, L. J., Davidowitz, G. & Nijhout, H. F. 35. Wullschleger, S., Loewith, R. & Hall, M. N. signaling pathway. Development 127, 3543–3551
The developmental and physiological basis of body TOR signaling in growth and metabolism. Cell 124, (2000).
size evolution in an insect. Proc. Biol. Sci. 268, 471–484 (2006). 58. De Jong, G. & Bochdanovits, Z. Latitudinal
1589–1593 (2001). 36. Colombani, J. et al. A nutrient sensor mechanism clines in Drosophila melanogaster: body size,
12. Ashburner, M. Drosophila, A Laboratory Handbook controls Drosophila growth. Cell 114, 739–749 (2003). allozyme frequencies, inversion frequencies, and the
(Cold Sping Harbor Press, Cold Spring Harbor, 1989). An interesting and creative study that implicates insulin-signalling pathway. J. Genet. 82, 207–223
13. Britton, J. S. & Edgar, B. A. Environmental control of the fat body as an important regulator of IIS (2003).
the cell cycle in Drosophila: nutrition activates mitotic signalling, and non-autonomous growth in 59. Edgar, B. & Nijhout, H. F. in Cell Growth: Control of Cell
and endoreplicative cells by distinct mechanisms. D. melanogaster. Size (eds Hall, M. N., Raff, M. & Thomas, G.) 23–83
Development 125, 2149–2158 (1998). 37. Radimerski, T. et al. dS6K regulated cell growth is (Cold Spring Harbor Lab. Press, Cold Spring Harbor,
14. Kim, S. K. & Rulifson, E. J. Conserved mechanisms of dPKB/dPI3K independent, but requires dPDK1. 2004).
glucose sensing and regulation by Drosophila corpora Nature Cell Biol. 4, 251–255 (2001). 60. Frei, C., Galloni, M., Hafen, E. & Edgar, B. A. The
cardiaca cells. Nature 431, 316–320 (2004). 38. Sarbassov, D. D., Guertin, D. A., Ali, S. M. & Sabatini, Drosophila mitochondrial ribosomal protein mRpL12
15. Lee, G. & Park, J. H. Hemolymph sugar homeostasis D. M. Phosphorylation and regulation of Akt/PKB by is required for Cyclin D/Cdk4-driven growth. EMBO J.
and starvation-induced hyperactivity affected by the rictor-mTOR complex. Science 307, 1098–1101 24, 623–634 (2005).
genetic manipulations of the adipokinetic hormone- (2005). 61. Grewal, S. S., Li, L., Orian, A., Eisenman, R. N. &
encoding gene in Drosophila melanogaster. 39. Puig, O. & Tjian, R. Transcriptional feedback control of Edgar, B. A. Myc-dependent regulation of ribosomal
Genetics 167, 311–323 (2004). insulin receptor by dFOXO/FOXO1. Genes Dev. 19, RNA synthesis during Drosophila development.
16. Broughton, S. J. et al. Longer lifespan, altered 2435–2446 (2005). Nature Cell Biol. 7, 295–302 (2005).
metabolism, and stress resistance in Drosophila from 40. Oldham, S., Bohni, R., Stocker, H., Brogiolo, W. & 62. Datar, S. A., Jacobs, H. W., de la Cruz, A. F.,
ablation of cells making insulin-like ligands. Proc. Natl Hafen, E. Genetic control of size in Drosophila. Philos. Lehner, C. F. & Edgar, B. A. The Drosophila cyclin
Acad. Sci. USA 102, 3105–3110 (2005). Trans. R. Soc. Lond. B Biol. Sci. 355, 945–952 D–Cdk4 complex promotes cellular growth. EMBO J.
17. Rulifson, E. J., Kim, S. K. & Nusse, R. Ablation of (2000). 19, 4543–4554 (2000).
insulin-producing neurons in flies: growth and diabetic 41. Masumura, M., Satake, S., Saegusa, H. & Mizoguchi, 63. Zwaan, B. J., Azevedo, R. B., James, A. C., Van’t Land, J.
phenotypes. Science 296, 1118–1120 (2002). A. Glucose stimulates the release of bombyxin, an & Partridge, L. Cellular basis of wing size variation in
18. Shingleton, A. W., Das, J., Vinicius, L. & Stern, D. L. insulin-related peptide of the silkworm Bombyx mori. Drosophila melanogaster: a comparison of latitudinal
The temporal requirements for insulin signaling during Gen. Comp. Endocrinol. 118, 393–399 (2000). clines on two continents. Heredity 84, 338–347
development in Drosophila. PLoS Biol. 3, e289 (2005). 42. Saucedo, L. J. et al. Rheb promotes cell growth as a (2000).
A thorough study of the function of component of the insulin/TOR signalling network. 64. Emlen, D. J., Szafran, Q., Corley, L. S. & Dworkin, I.
the roles of InR in growth control in Nature Cell Biol. 5, 566–571 (2003). Insulin signaling and limb-patterning: candidate
D. melanogaster, delineating its temporal 43. Rusten, T. E. et al. Programmed autophagy in the pathways for the origin and evolutionary diversification
requirements and identifying its role in allometric Drosophila fat body is induced by ecdysone through of beetle ‘horns’. Heredity 97, 179–191 (2006).
growth. regulation of the PI3K pathway. Dev. Cell 7, 179–192 65. Bryant, P. J. & Levinson, P. Intrinsic growth control in
19. Wu, Q. & Brown, M. R. Signaling and function of (2004). the imaginal primordia of Drosophila, and the
insulin-like peptides in insects. Annu. Rev. Entomol. 44. Scott, R. C., Schuldiner, O. & Neufeld, T. P. Role and autonomous action of a lethal mutation causing
51, 1–24 (2006). regulation of starvation-induced autophagy in the overgrowth. Dev. Biol. 107, 355–363 (1985).
20. Efstradiatis, A. Genetics of mouse growth. Int. J. Dev. Drosophila fat body. Dev. Cell 7, 167–178 (2004). 66. Zitnan, D., Sehnal, F. & Bryant, P. J. Neurons
Biol. 42, 955–976 (1998). 45. Truman, J. W. et al. Juvenile hormone is required to producing specific neuropeptides in the central
21. Baserga, R. in Cell Growth (eds Hall, M. N., Raff, M. & couple imaginal disc formation with nutrition in nervous system of normal and pupariation-delayed
Thomas, G.) 235–262 (Cold Spring Harbor Press, insects. Science 312, 1385–1388 (2006). Drosophila. Dev. Biol. 156, 117–135 (1993).
Cold Spring Harbor, 2004). A provocative and sometimes perplexing study that 67. Sehnal, F. & Bryant, P. J. Delayed pupariation in
22. McGuire, S. E., Roman, G. & Davis, R. L. details the relationship between juvenile hormones Drosophila imaginal disc overgrowth mutants is
Gene expression systems in Drosophila: a synthesis of and nutrition as regulators of growth in the late- associated with reduced ecdysteroid titer. J. Insect
time and space. Trends Genet. 20, 384–391 (2004). forming eye and wing discs of M. sexta. Physiol. 12, 1051–1059. (1993).

NATURE REVIEWS | GENETICS VOLUME 7 | DECEMBER 2006 | 915


© 2006 Nature Publishing Group
REVIEWS

68. Brumby, A. M. & Richardson, H. E. Using Drosophila 79. Davidowitz, G. & Nijhout, H. F. The physiological basis 90. Truman, J. W. & Riddiford, L. M. Physiology of insect
melanogaster to map human cancer pathways. of reaction norms: the interaction among growth rate, rhythms. 3. The temporal organization of the
Nature Rev. Cancer 5, 626–639 (2005). the duration of growth and body size. Integr. Comp. endocrine events underlying pupation of the tobacco
69. Shearn, A., Rice, T., Garen, A. & Gehring, W. Biol. 144, 443–449 (2004). hornworm. J. Exp. Biol. 60, 371–382 (1974).
Imaginal disc abnormalities in lethal mutants of 80. Atkinson, D. Temperature and organism size — 91. Parvy, J. P. et al. A role for βFTZ-F1 in regulating
Drosophila. Proc. Natl Acad. Sci. USA 68, a biological law for ectotherms? Adv. Ecol. Res. 25, ecdysteroid titers during post-embryonic development
2594–2598 (1971). 1–58 (1994). in Drosophila melanogaster. Dev. Biol. 282, 84–94
70. Shearn, A. & Garen, A. Genetic control of imaginal disc 81. French, V., Feast, M. & Partridge, L. Body size and cell (2005).
development in Drosophila. Proc. Natl Acad. Sci. USA size in Drosophila: the developmental response to 92. James, A. C., Azevedo, R. B. & Partridge, L. Cellular basis
71, 1393–1397 (1974). temperature. J. Insect Physiol. 44, 1081–1089 (1998). and developmental timing in a size cline of Drosophila
71. Szabad, J. & Bryant, P. J. The mode of action of 82. Azevedo, R. B. R., French, V. & Partridge, L. melanogaster. Genetics 140, 659–666 (1995).
‘‘discless’’ mutations in Drosophila melanogaster. Temperature modulates epidermal cell size in 93. Partridge, L., Barrie, B. Fowler K. & French, V.
Dev. Biol. 93, 240–256 (1982). Drosophila melanogaster. J. Insect Physiol. 48, Evolution and development of body size and cell size in
72. Browder, M. H., D’Amico, L. J. & Nijhout, H. F. 231–237 (2002). Drosophila melanogaster in response to temperature.
The role of low levels of juvenile hormone esterase in 83. Puig, O., Marr, M. T., Ruhf, M. L. & Tjian, R. Control of Evolution 48, 1269–1276 (1994).
the metamorphosis of Manduca sexta. J. Insect Sci. 1, cell number by Drosophila FOXO: downstream and 94. Partridge, L., Langelan, R., Fowler, K., Zwaan, B. &
11 (2001). feedback regulation of the insulin receptor pathway. French, V. Correlated responses to selection on body
73. King-Jones, K., Charles, J. P., Lam, G. & Thummel, C. S. Genes Dev. 17, 2006–2020 (2003). size in Drosophila melanogaster. Genet. Res. 74,
The ecdysone-induced DHR4 orphan nuclear receptor 84. Hennig, K. M., Colombani, J. & Neufeld, T. P. 43–54 (1999).
coordinates growth and maturation in Drosophila. Cell TOR coordinates bulk and targeted endocytosis in 95. Gockel, J., Kennington, W. J., Hoffmann, A., Goldstein,
121, 773–784 (2005). the Drosophila melanogaster fat body to regulate cell D. B. & Partridge, L. Nonclinality of molecular
74. Calboli, F. C., Gilchrist, G. W. & Partridge, L. Different growth. J. Cell Biol. 173, 963–974 (2006). variation implicates selection in maintaining a
cell size and cell number contribution in two newly 85. MacWhinnie, S. G. et al. The role of nutrition in morphological cline of Drosophila melanogaster.
established and one ancient body size cline of creation of the eye imaginal disc and initiation of Genetics 158, 319–323 (2001).
Drosophila subobscura. Evolution 57, 566–573 metamorphosis in Manduca sexta. Dev. Biol. 285,
(2003). 285–297 (2005). Competing interests statement
75. Bryant, P. J. & Levinson, P. Intrinsic growth 86. Flatt, T., Tu, M. P. & Tatar, M. Hormonal pleiotropy The author declares no competing financial interests.
control in the imaginal primordia of Drosophila, and the juvenile hormone regulation of Drosophila
and the autonomous action of a lethal mutation development and life history. BioEssays 27,
causing overgrowth. Dev. Biol. 107, 355–363 999–1010 (2005). DATABASES
(1985). 87. Yamanaka, N. et al. From the cover: regulation of The following terms in this article are linked online to:
76. Truman, J. W. & Bate, M. Spatial and temporal insect steroid hormone biosynthesis by innervating Entrez Gene: http://www.ncbi.nlm.nih.gov/entrez/query.
patterns of neurogenesis in the central nervous peptidergic neurons. Proc. Natl Acad. Sci. USA 103, fcgi?db=gene
system of Drosophila melanogaster. Dev. Biol. 125, 8622–8627 (2006). AKT | chico | 4EBP | FOXO | HH | InR | Notch | PDK1 | PTEN |
145–157 (1988). 88. Shafiei, M., Moczek, A. P. & Nijhout, H. F. Food S6K | TIF-IA | TOR | WG
77. Sears, K. E., Behringer, R. R., Rasweiler, J. J. T. & availability controls the onset of metamorphosis in the UniProtKB: http://ca.expasy.org/sprot
Niswander, L. A. Development of bat flight: dung beetle Onthophagus taurus (Coleoptera: Myc
morphologic and molecular evolution of bat wing Scarabeidae). Physiol. Entomol. 26, 173–180 (2001).
digits. Proc. Natl Acad. Sci. USA 103, 6581–6586 89. Riddiford, L. M., Hiruma, K., Zhou, X. & Nelson, C. A. FURTHER INFORMATION
(2006). Insights into the molecular basis of the hormonal Edgar laboratory homepage:
78. Emlen, D. J. & Nijhout, H. F. The development and control of molting and metamorphosis from Manduca http://www.fhcrc.org/science/labs/edgar
evolution of exaggerated morphologies in insects. sexta and Drosophila melanogaster. Insect Biochem. Access to this links box is available online.
Annu. Rev. Entomol. 45, 661–708 (2000). Mol. Biol. 33, 1327–1338 (2003).

916 | DECEMBER 2006 | VOLUME 7 www.nature.com/reviews/genetics


© 2006 Nature Publishing Group

You might also like