ArticlePDF Available

Drosophila melanogaster as a Versatile Model Organism in Food and Nutrition Research

Authors:

Abstract and Figures

Drosophila melanogaster has been widely used in the biological sciences as a model organism. Drosophila has a relatively short life span of 60–80 days, which makes it attractive for life span studies. Moreover, approximately 60% of the fruit fly genes are orthologs to mammals. Thus, metabolic and signal transduction pathways are highly conserved. Maintenance and reproduction of Drosophila do not require sophisticated equipment and are rather cheap. Furthermore, there are fewer ethical issues involved in experimental Drosophila research compared with studies in laboratory rodents, such as rats and mice. Drosophila is increasingly recognized as a model organism in food and nutrition research. Drosophila is often fed complex solid diets based on yeast, corn, and agar. There are also so-called holidic diets available that are defined in terms of their amino acid, fatty acid, carbohydrate, vitamin, mineral, and trace element compositions. Feed intake, body composition, locomotor activity, intestinal barrier function, microbiota, cognition, fertility, aging, and life span can be systematically determined in Drosophila in response to dietary factors. Furthermore, diet-induced pathophysiological mechanisms including inflammation and stress responses may be evaluated in the fly under defined experimental conditions. Here, we critically evaluate Drosophila melanogaster as a versatile model organism in experimental food and nutrition research, review the corresponding data in the literature, and make suggestions for future directions of research.
Content may be subject to copyright.
1
Drosophila melanogaster as a versatile model organism in food and 1
nutrition research 2
Stefanie Staatsa*, Kai Lüersena, Anika E. Wagnerb, Gerald Rimbacha 3
a Institute of Human Nutrition and Food Science, Christian-Albrechts-University Kiel, Hermann-4
Rodewald-Strasse 6, D-24118 Kiel, Germany 5
b Institute of Nutritional Medicine, University of Lübeck, Ratzeburger Allee 160, D-23538 Lübeck, 6
Germany 7
* Corresponding author: Tel +49 431 880 5313. Fax +49 431 880 2628. E-mail address: 8
staats@foodsci.uni-kiel.de 9
2
Abstract 10
Drosophila melanogaster has been widely used in the biological sciences as a model organism. 11
Drosophila has a relatively short life span of 60 to 80 days, which makes it attractive for life span studies. 12
Moreover, approximately 60% of the fruit fly genes are orthologs to mammals. Thus, metabolic and 13
signal transduction pathways are highly conserved. Maintenance and reproduction of Drosophila do not 14
require sophisticated equipment and are rather cheap. Furthermore, there are fewer ethical issues 15
involved in experimental Drosophila research compared with studies in laboratory rodents, such as rats 16
and mice. Drosophila is increasingly recognized as a model organism in food and nutrition research. 17
Drosophila is often fed complex solid diets based on yeast, corn and agar. There are also so-called 18
holidic diets available that are defined in terms of their amino acid, fatty acid, carbohydrate, vitamin, 19
mineral and trace element compositions. Feed intake, body composition, locomotor activity, intestinal 20
barrier function, microbiota, cognition, fertility, ageing and life span can be systematically determined 21
in Drosophila in response to dietary factors. Furthermore, diet-induced pathophysiological mechanisms 22
including inflammation and stress responses may be evaluated in the fly under defined experimental 23
conditions. Here, we critically evaluate Drosophila melanogaster as a versatile model organism in 24
experimental food and nutrition research, review the corresponding data in the literature and make 25
suggestions for future directions of research. 26
Key words: Drosophila melanogaster, model organism, nutrition, longevity, metabolism
27
3
Introduction 28
Drosophila melanogaster is a proven model organism in genetic research. The fruit fly has been further 29
established as an emerging and valuable model in experimental food and nutrition research in the past 30
few decades. D. melanogaster is suitable to be used in nutritional intervention studies as it exhibits many 31
similarities with mammalian species. Although the insect body plan is simpler than that of mammals, 32
the anatomy of fruit flies includes organ systems with equivalent functions of the mammalian heart, 33
lung, kidney, liver and gonads. Moreover, the fruit fly holds a complex and dynamic gut exhibiting a 34
similar structure and organization of the mammalian gut. Also, the tissues, physiology and anatomy of 35
mammalian and D. melanogaster intestines exhibit similar properties. Moreover, the fruit fly possesses 36
a central and peripheral nervous system, produces gastrointestinal and sex hormones like insulin-like 37
peptides, juvenile hormone and ecdysone that affect the fly’s metabolism and development, as well as 38
occupies genes and proteins displaying 50-60% orthology to mammalian ones. Within this review article 39
we describe fly anatomy and physiology including nutrient sensing and endocrine signaling. Since feed 40
intake and the composition of experimental diets are important determinants in nutritional studies, we 41
introduce different methods as far as the quantification of feed intake and also the examination of food 42
preferences are concerned. Furthermore, we discuss the pros and cons of complex versus so-called 43
holidic diets. We then compile different Drosophila studies concerning the impact of diet on various 44
outcome measurements including life span and longevity-associated gene expression, locomotor 45
activity, intestinal barrier function and gut microbiota as well as fertility. Finally, we make suggestions 46
in terms of future directions of research applying D. melanogaster as a model organism in food and 47
nutritional sciences. 48
Fly anatomy 49
a. Morphology and gut anatomy 50
Drosophila melanogaster is a holometabolous insect. Hence, its life cycle encompasses the four 51
developmental stages embryo, larvae (first, second and third instar larvae separated by molts), pupa and 52
adult, of which only the larvae and the adult flies are active feeders. In accordance with a typical insect 53
4
morphology, the segmented body of an adult D. melanogaster can be divided into three parts, the head, 54
the thorax and the abdomen. The head carries sucking mouthparts and sensory organs, including a pairs 55
of antennae and compound eyes. Each of the three thoracic segments (pro-, meso- and metathorax) bears 56
a pair of limbs. In addition, a pair of wings is attached to the mesothorax and a pair of halters is attached 57
to the metathoracic segment. The abdominal segments lack extremities but contain the male and female 58
genitalia (Fig. 1A) 1. The worm-like body of larvae is also segmented but lacks compound eyes, limbs 59
and wings. Adult flies ingest mainly liquid food via their proboscis, whereas the mouth hooks of larvae 60
allows the ingestion of solid food 2. 61
Although the insect body plan is simpler than that of mammals, the anatomy of the fruit fly includes 62
organ systems with functions equivalent to the mammalian brain and peripheral nervous system, heart, 63
lung (trachea system in the fly), kidney (Malpighian tubules in the fly), liver (fat body in the fly), gut 64
and gonads 3-7. For instance, like the mammalian gastro-intestinal system, the digestive tract of the fruit 65
fly is responsible for the digestion and uptake of nutrients. The gut tube that traverses the entire body of 66
adult animals is lined by a simple epithelium of columnar or cuboidal cells called enterocytes that secrete 67
digestive enzymes and absorb nutrients. As in mammals, the basal side of this epithelial monolayer is 68
aligned to the basement membrane, an extracellular collagenous matrix. Visceral muscles that surround 69
the epithelial tube elicit peristaltic movement. They are innervated by the central nervous system and 70
oxygenated by trachea 2, 8. Associated enteroendocrine cells are responsible for humoral signaling and 71
stem cells enable regeneration processes e.g. the Drosophila midgut epithelium is completely renewed 72
every 1-2 weeks 9, 10. 73
The alimentary canal of Drosophila consists of three main parts, the foregut, the midgut and the hindgut. 74
As in mammals, the different parts of the Drosophila digestive tract are highly specialized in anatomy, 75
organization and function 2, 8. The foregut starts with the oral cavity (pharynx) followed by the 76
esophagus, which is connected to the crop, a food storage organ. The foregut and midgut are separated 77
by a valve-like organ, the cardia, whose foregut portion is the proventriculus. The cardia secretes the 78
peritrophic matrix, a non-cellular semi-permeable structure that is composed of chitin and glycoproteins, 79
and surrounds the food bolus in the midgut. Analogous to the role of mucous secretions of the vertebrate 80
5
digestive tract, the peritrophic matrix forms a physical barrier that is thought to protect the Drosophila 81
midgut epithelium from harmful particles and pathogens 11. According to morphometric parameters the 82
midgut consists of six regions. However, molecular and transcript analyses revealed that the midgut 83
region can be further subdivided into 14 functional zones along its anterior–posterior axis 12. In 84
particular, a stomach-like acidic compartment of the Drosophila midgut has been determined. Here, 85
copper cells secret acid in a manner similar to mammalian gastric parietal cells, which leads to a local 86
pH < 3 8, 9. Malpighian tubules, which play a role analogous to the kidneys in vertebrates, are connected 87
to the gut at the midgut-hindgut junction. The hindgut functions in water and ion reabsorption prior to 88
fecal excretion, similar to the mammalian large intestine. The hindgut ends in the anus, which is located 89
at the last abdominal segment 2, 8. 90
The overall organization of the larval gastrointestinal tract is similar to that of the adult fly. However, 91
larvae lack a crop but contain four gastric ceca, blind-ending pouches attached to the anterior midgut, 92
that do not exist in the adult alimentary tract 2, 8. Comparable to the situation in adults, the larval midgut 93
can be functionally subdivided into specific segments 8, 13, 14. During metamorphosis, the larval gut is 94
replaced by a de novo generated adult intestinal tract 2. 95
b. Size, body weight, body composition 96
Size is determined in terms of thorax length in Drosophila and is affected by ambient temperature 15, 16 97
and by genetic 17 and dietary factors 18 which affect mating behavior and survival 19. Nutrition may affect 98
larval development and the adult body weight of fruit flies either by modifying the fly’s metabolism or 99
by direct caloric restriction due to a reduced feed intake. Effects of nutritional intervention on body 100
weight development can be easily assessed by live weighing of a defined pool of flies 20. Male flies 101
weigh approximately 700 µg while female flies weigh 1000-1200 µg. Body composition can be defined 102
by means of lean body mass, fat and protein content and depends on the delicate adjustment of catabolic 103
and anabolic pathways in the fruit fly. Thus, body composition can be modified by nutritional 104
compounds and dietary supplements that affect the metabolism of Drosophila. Changes in fruit fly body 105
composition can be determined by measuring whole body protein and triglyceride levels following fly 106
homogenization and lysis using commercially available kits e.g., the bicinchoninic acid colorimetric 107
6
assay and the lipase/glycerol kinase/peroxidase-based colorimetric assay 20. Similarly, glucose and 108
trehalose levels, the most important energy resources and storage carbohydrates in Drosophila 21, 22, are 109
quantifiable using a commercial glucose oxidase/peroxidase-based colorimetric assay optionally 110
combined with trehalose digestion 20, 22. Importantly, glucose and triglyceride levels determined in 111
Drosophila should be normalized to live body weight rather than to total protein content. In in vitro 112
studies, protein normalization is often used as an internal control, but in the fruit fly, nutritional 113
intervention may affect the total protein content, thus normalization to the protein content may lead to 114
false results. In addition, the water content can be determined by calculating the difference between a 115
fly’s live and dry (following drying with the help of heat or CaCl2) weights 23-25. Both body weight and 116
body composition may vary dependent on the fly strain and the age of the flies (Abb. 1B). The 117
application of the abovementioned methods and kits to determine macronutrients in the fruit fly itself is 118
functional and allows the analysis of small sample pools of flies or isolated tissue and organs with an 119
adequate repetition number to minimize quantitation errors between different measurement days. In 120
order to determine macronutrient composition in the fly feed, standard AOAC analytical methods are 121
recommended. 122
c. Enzymatic machinery 123
The digestion and absorption of food occurs predominantly in the midgut of fruit flies. The distinct 124
enzymatic infrastructure of the different midgut zones is an important prerequisite for an ordered 125
digestion and absorption process within the Drosophila gut lumen 2, 8. Like mammals, dietary 126
macronutrients have to be broken down prior to absorption by the enterocytes of D. melanogaster. 127
According to an in silico search by Lemaitre and Miguel-Aliaga (2013), 349 putative digestive enzymes 128
have been predicted in the fruit fly genome 2. The vast majority of the hydrolases involved in these 129
digestive processes are homolog to the respective mammalian enzymes found in the intestinal tract. As 130
in humans, dietary polysaccharides, such as starch, are broken down, usually into monosaccharides, in 131
the Drosophila gut lumen by the consecutive action of alpha-amylases (Amy-p, Amy-d, and Amyrel) 132
and several alpha-glucosidases (eight putative alpha-glucosidase genes have been identified, e.g. 133
glucosidase 1, glucosidase 2 α and β) 26-28. Up to date, there are no indications for the subsistence of 134
7
enzymes with dietary beta-glucosidase activity in the fruit fly. Nevertheless, D. melanogaster exhibits 135
the lysosomal-expressed glucocerebrosidase 1b (Gba1b) cleaving glucosylceramide and playing a major 136
role in Parkinson’s disease 29. Dietary lipids, such as triacylglycerols, phospholipids and other 137
acylglycerols, are digested by the action of neutral lipases that are secreted into the midgut lumen, similar 138
to their human counterparts 30-32. Moreover, the fruit fly is equipped with a broad range of endo- and 139
exopeptidases, including e.g., trypsin that cleave dietary proteins and peptides into dipeptides and amino 140
acids 2, 33. In humans, the end products of luminal digestion (monosaccharides and dipeptides/amino 141
acids) are then taken up by specific transport systems located in the apical membrane of the small 142
intestine. Given that many genes homologous to known mammalian transporter family members are 143
present in the genome of D. melanogaster, one can assume similar uptake mechanisms for the midgut 144
epithelium in fruit flies 2. Nevertheless, it is remarkable that these transport processes and the 145
participating transporter proteins are not yet well characterized. An exception is the dipeptide transporter 146
Opt1, a homolog of the human SLC15 dipeptide transporters pepT1 and pepT2. Drosophila Opt1 is 147
expressed on the apical membrane of the midgut epithelium and exhibits proton-dependent, high-affinity 148
dipeptide transport activity 34. To be dispersed within the entire body, the absorbed nutrients have to be 149
released into the fly’s open circulatory system, which is filled with hemolymph, a blood-like fluid. 150
Similar to humans, the transfer of lipids across the gut epithelium and their further transport in the 151
hemolymph requires lipoproteins. Following absorption into midgut enterocytes, dietary fatty acids are 152
incorporated into diacylglycerols, which together with dietary sterols, are bound to the apolipoprotein B 153
(ApoB)-family lipoprotein lipid transfer particle (LTP). For hemolymph transport, the lipids are shifted 154
onto lipophorin (Lpp), another ApoB-family lipoprotein 35. The Drosophila intestinal tract is also a good 155
model to study the transporter-mediated uptake processes of micronutrients such as zinc (via the 156
SLC39A zinc transporters dZIP) and copper (via the copper transporters CtrA and CtrB) 36, 37. 157
d. Nutrient sensing and endocrine signaling 158
Similar to mammalian species, D. melanogaster has a complex neuroendocrine system that produces 159
peptide and steroid hormones, such as insulin-like peptides, juvenile hormone and ecdysone 38-42, to 160
control the fly’s metabolism and development. The actual levels of certain nutrients affect the release of 161
8
hormones and control nutrient- and energy-sensing signaling pathways, which then modulate 162
metabolism to maintain homeostasis at the cellular and organismic level. In humans, the sensing of 163
ingested food starts with the enteroendocrine system in the intestinal tract, which leads to the secretion 164
of gastrointestinal peptide hormones such as cholecystokinin (CCK), glucagon-like peptide (GLP1), 165
peptide YY and glucose-dependent insulinotropic peptide (GIP) 43. These hormones affect peristaltic 166
movement, digestive enzyme secretion, glucose homeostasis and appetite. The enteroendocrine cells of 167
the D. melanogaster midgut have been shown to express a battery of potential regulatory peptides. It is 168
very likely that most of these peptides are involved in nutrient sensing, the control of gut function and 169
metabolic homeostasis 44, 45. However, up to now a physiological function could be ascribed to only a 170
limited number of these peptides. For example, activin-β (Actβ) expression is upregulated in 171
enteroendocrine cells by a chronically high sugar diet and counteracts hyperglycemia by eliciting the 172
action of adipokinetic hormone (AKH, a glucagon-like peptide; see below) in the fat body 46, 47. 173
The best-known example of nutrient-induced hormone regulation is the impact of the serum glucose 174
level on the release of the antagonistic peptide hormones insulin and glucagon from mammalian 175
pancreatic β- and α-cells, respectively. Insulin is an indicator for high glucose and high energy levels, 176
whereas its counterpart, glucagon, is secreted under low glucose and low energy conditions. The 177
interplay between insulin and glucagon is also crucial for systemic glucose homeostasis in D. 178
melanogaster. The fly possesses eight homologous insulin-like peptides (dILP1-8) and the glucagon-179
like peptide AKH. A glucose-rich diet induces the secretion of dILP2, dILP3 and dILP5 that are mainly 180
produced in 14 β-cell-like insulin producing cells (IPCs) located in the central nervous system of the fly 181
48, 49. Similar to the mechanism reported for mammalian β-cells, dILP release from IPCs is regulated 182
autonomously by glucose sensing performed by glucose transporters, KATP channels, voltage sensitive 183
Ca2+ channels and dietary amino acids, in particular leucine 48, 50. Moreover, as in mammals, several 184
neuronal and humoral inter-organ communication networks have been identified that link the systemic 185
nutrition and/or energy status sensed in peripheral organs to IPC activity 48, 51-54. The fat body, the 186
functional analog of the mammalian liver/adipose tissue, and the intestine are metabolic key integrators 187
in these processes. The Sirt-UPD2-IIS axis, e.g., which connects the nutrition/energy status of the fat 188
9
body to IPC activity has been deciphered 51. The evolutionarily conserved NAD+-dependent metabolic 189
sensor Sir2/Sirt1 stimulates the release of UPD2, the functional homolog of mammalian leptin, from the 190
fat body into the hemolymph, specifically in response to glucose or lipid feeding. As a result, the 191
JAK/STAT ligand UPD2 increases dILP release from IPCs most probably by inhibition of GABAergic 192
brain neurons that tonically inhibit IPC 48, 51. 193
As in mammals, binding of dILP to the insulin/IGF-like tyrosine kinase receptor (INR) activates 194
phosphoinositide 3-kinase (PI3K)-dependent signaling pathways via the Drosophila insulin receptor 195
substrate (IRS) Chico, thereby promoting the activation of the serine/threonine-specific protein kinase 196
AKT. Owing to its phosphorylating activity, AKT prevents the translocation of the forkhead 197
transcription factor FOXO from the cytosol into the nucleus. In target tissues, dILP signaling has 198
anabolic effects, increases nutrient storages and supports growth 48, 49, 55. 199
The glucagon-like peptide AKH is synthesized in α-cell-like corpus cardiacum (CC) cells that are part 200
of the ring gland directly connected to the fly’s aorta. Similar to human α-cells, CC cells respond to low 201
sugar levels in the hemolymph by activating the internal low energy status sensor adenosine 202
monophosphate-activated kinase (AMPK) which results in Ca2+ dependent AKH release. In accordance 203
with its antagonistic function to insulin, AKH elicits the mobilization of glycogen and lipid stores in 204
target tissues via binding to the G protein-coupled AKH receptor 56. 205
In addition to this insulin mechanism, the abundance of certain amino acids is also sensed by the target 206
of rapamycin (TOR) pathway. As in humans, high amino acid levels activate the Drosophila dTOR 207
pathway leading to the phosphorylation of kinase S6K and eIF-4E binding protein (4E-BP), which 208
promote the initiation of translation and elongation as well as ribosome biogenesis (Oldham, 2011). 209
Similar to mammals, dTOR is also controlled by cellular and systemic nutrient statuses. A high 210
intracellular adenosine monophosphate ÷ adenosine triphosphate ratio inhibits dTOR activity through 211
AMPK. Indicative of a high systemic nutrient and energy status, AKT, a key integration factor and 212
downstream target of IIS signaling (see above), directly inhibits tuberous sclerosis tumor suppressor 213
(TSC1/2), which in turn inhibits the small GTPase Ras homolog enriched in brain (Rheb). Since Rheb 214
10
functions as a negative dTOR regulator, IIS/AKT promotes dTOR signaling 48, 49. Hence, similar to 215
humans, in fruit flies, the IIS and dTOR pathways converge to regulate protein synthesis and growth. 216
Measurement of feed intake and composition of experimental diets 217
a. Quantification of food ingestion and examination of marked feed preferences 218
Administration of various dietary factors such as secondary plant compounds to the Drosophila standard 219
medium may affect the taste of the food due to sweetness, bitterness or saltiness and thus might result in 220
reduced food uptake by the supplemented flies. As calorie restriction extends the life span of the fruit 221
fly 57-59, alters body weight 57, may affect molecular signaling pathways and locomotor activity 60 and 222
might mask plant bioactive-dependent effects on the fly’s metabolism, equal food uptake has to be 223
ascertained by determining the quantity of ingested food. Moreover, the general acceptance of 224
administered plant metabolites can be evaluated through the use of choice assays. 225
Ingestion quantity. To quantify food uptake in Drosophila, e.g., to exclude dietary restriction-226
dependent effects following administration of an ill-tasting secondary plant compound, both 227
administration of food colorings, radioisotope-labelling with 32P, observation of proboscis extension 228
(PE) and application of the capillary feeding (CAFE) assay are applicable. 229
Estimating food uptake by dye usage is easy and inexpensive while it can be rapidly assessed by visual 230
inspection of the flies. FCF Brilliant Blue and Sulforhodamine B (Fig. 2, left panel) represent suitable 231
food dyes that are added to the fly’s standard food medium at concentrations of 2.5% and 0.2% (w/v) 232
and do not affect food intake per se 58, 61, 62. Flies that ingest the dyes display colored thoraxes and 233
abdomens, particularly tinted fat bodies (Fig. 2, right panel). Food uptake can be calculated by scoring 234
the intensity of body coloring following visual evaluation and by using photometric and fluorometric 235
measurements of the processed flies at dye-specific wavelengths adjusted to a standard curve. 236
Quantification of food uptake by a dye-based method is rather imprecise and challenging in term of its 237
reproducibility 62. However, dye-based methods can be further improved by using radioactive tracers, as 238
radioisotopes are effectively incorporated and retained in the flies bodies. Thus, administration of 32P at 239
11
concentrations of 0.5–4 μCi/ml [α-32P] dCTP allows for consistent and time-dependent measurement 240
of food uptake while quantification is carried out by a scintillation counter 59, 62-64. 241
The CAFE assay is more complex but more precise, highly reproducible and allows for the continuous 242
real-time quantitation of food ingestion in individual fruit flies. This method is based on the feeding of 243
a liquid food preparation instead of the solid fly medium 55. Flies are maintained in experimental vials 244
closed with a specific lid that is pierced with a microcapillary containing the liquid food. Food uptake is 245
calculated with the help of a graduated glass microcapillary with a holding capacity of 5 µl 65. Food 246
color is dispensable but can be used for better visualization 65. 247
Another possibility to quantify food uptake in Drosophila is to observe the proboscis extension (PE) of 248
flies maintained on the experimental food medium. The proboscis extension reflex is a congenital 249
behavior of wild-type flies ensuring food uptake 66, 67. Food uptake is defined as the event when a fly 250
extends its proboscis, touches the medium surface and performs a bobbing motion. Events are scored by 251
an observer repeatedly monitoring the flies for 3 sec every 2-5 min for a total of 90 min 68. However, the 252
PE assay is not recommended as the sole feeding assay as it hardly detects reproducible differences and 253
shows considerable variability in feed intake in the fruit fly thus increasing the risk of obtaining false 254
negative results 62. 255
Similarly, food intake can be determined in Drosophila larvae. Preferably, food-dye based methods 256
using FCF Brilliant blue (2.5% w/v) are used. Following a defined feeding period, food intake is 257
photometrically quantified in tinted larvae following homogenization of whole larvae or isolated guts 69,
258
70. 259
Food choice. In addition to the abovementioned food intake assays, there are some methods available 260
that allow for the examination of food preferences. The fly’s proclivity for a specific macronutrient 261
composition and aversion to some secondary plant compounds due to its bitterness can be evaluated 262
using the two-choice preference test. Therefore, two different liquid food preparations containing e.g., 263
specific plant bioactives or micronutrients, are stained with red and blue food dyes, respectively. Food 264
is applied to filter discs, put into an enclosed box and the flies are allowed to ingest for a defined interval. 265
The food dyes accumulate in the flies as they eat and food uptake can be evaluated visually (red-, blue- 266
12
and purple-colored flies) and photometrically or fluorometrically 67. The choice assay can also be 267
performed with the help of the CAFE assay setup while providing different food preparations in a 268
number of capillaries 65. Further, a modified version of the PE assay, the manual feeding assay (MAFE) 269
71 combined with the proboscis extension reflex assay (PER) 66, 67 can be applied to test Drosophila for 270
its ingestion readiness of selected nutrients and substances. Therefore, fixed flies manually receive a 271
liquid food preparation repeatedly delivered by a pipette tip or a fine graduated capillary. A feeding (PE) 272
event is counted if flies fully extend their proboscises and start drinking the liquid food. Food preference 273
and food quantity can be simultaneously determined by calculating the number of PE events and by 274
reading out the ingestion volume from the capillary 71. 275
Applying the FRAPPÉ assay is a specific opportunity to evaluate ethanol consumption in the fruit fly. 276
This assay allows disclosing alcohol preferences in Drosophila by fast and high throughput 277
measurements of consumption in individual ethanol-primed flies using a fluorescence plate reader. 278
Different liquid food preparations are stained with fluorescent dyes, such as rhodamine B and 279
fluorescein, and the flies are allowed to ingest the food for a defined period of time. The intensity of the 280
dye is correlated to the amount of food ingested and can be fluorometrically determined in whole flies 281
or whole body homogenates using a microplate reader 72. 282
In Drosophila larvae food preferences can be similarly estimated performing choice assays with 283
supplemented agarose-filled petri dishes, e.g., with sweet or bitter tasting macro- and micronutrients and 284
secondary plant compounds. Larvae in the third instar feeding-stage are washed and put on the prepared 285
agarose plates prior larvae counting following a defined time lapse 73-76. Similarly, petri dishes 286
containing filter paper soaked with the experimental liquid diet can be applied 69. The distribution of 287
larvae then points to the preferred food. 288
b. Complex versus chemically-defined diets 289
There are different methods available for the preparation of solid or semi-solid media for optimal 290
Drosophila larval growth, maintenance of adults and the performance of nutrition experiments. Most 291
common, solid media are prepared with varying amounts of sugars, agar and yeast serving as the protein 292
and micronutrient source, while sometimes corn meal is also added and inactive dry yeast is replaced by 293
13
brewer’s yeast 20, 77, 78. Although food ingredients are specifically manufactured for Drosophila 294
husbandry, these complex media do not have defined protein, vitamin and mineral concentrations. Even 295
more undefined media, based on bananas, jaggery, yeast and barley are occasionally used 79, 80. However, 296
many studies investigating the effect of nutritional compounds and plant metabolites on life span and 297
health performance parameters of the fruit fly use a modified Caltech medium 81 consisting of 5-10.5% 298
sucrose, 2.1-5% yeast, 8.6-10.5% corn meal and 0.5-1.3% agar 20, 78, 82-85. Accordingly, a sugar-yeast 299
medium of 10% sucrose, 10% yeast and 2% agar is commonly used 86, 87. Methylparaben and propionic 300
acid are commonly used as preservatives in Drosophila food preparations. The advantage of all of the 301
abovementioned media is that the preparation is simple, provides flies with all the essential nutrients and 302
reveals supplement-dependent effects on fly development and growth parameters if they have a 303
significant impact. In addition to ambient temperature and virginity 88, nutrient density might affect the 304
fly’s metabolism 88-91 and may alter its response to nutritional interventions leading to false-positive 305
amplification or false-negative attenuation of detectable effects 85. In this particular case, the application 306
of a chemically defined food preparation, a holidic medium, is applicable. Piper et al. introduced this 307
fully synthetic diet to the Drosophila experimental portfolio in 2014 92. In addition to sucrose and agar 308
that ensure the solidity of the medium, only isolated dietary factors are added. Therefore, defined 309
concentrations of cholesterol, minerals (Na, K, Ca, Cl, Cu, Fe, Mg, Mn, Zn and SO4, CO3, PO4), essential 310
and non-essential amino acids (isoleucine, leucine, arginine, histidine, lysine, phenylalanine, threonine, 311
tryptophan, valine and tyrosine, alanine, asparagine, aspartic acid, cysteine, glutamine, glycine, proline, 312
serine), glutamate, vitamins (B1, B2, B6, biotin, folic acid, niacin, Ca-pantothenate), choline, myo-313
inositol, inosine and uridine are added. The pH value is adjusted using a buffer. There are additional 314
recipes available for the preparation of chemically defined media that represent modifications of the 315
holidic diet. Compared with holidic medium, 400 Kcal/L chemically defined food (CDF400K) 316
additionally contains vitamins A, E, D3 and K as well as the metal ion Cr and lactose, glucose and 317
trehalose sugars 93. In contrast, Sang’s medium contains casein instead of isolated amino acids and is 318
supplemented with fructose and lecithin but lacks Cu, Zn, myo-inositol, inosine and uridine compared 319
with the holidic diet 94. Importantly, varying food preparations differ in their protein, carbohydrate, lipid, 320
14
vitamin, mineral and gelling agent content and, therefore, affect larval development, time lapse until 321
adult eclosion, maximum life span and fecundity differently 92, 95. For instance, a reduction or depletion 322
of sterols and single amino acids or vitamins in the diet markedly reduces fertility of the flies 92 while 323
increasing the carbohydrate content may significantly prolong life span 95. Moreover, considering the 324
impact of the nutritional geometry on life span and fitness parameters, not only micronutrients and 325
secondary plant compounds but also changes in macronutrient availability and energy content of the 326
food may determine the health and survival of the fly 59, 96-98. There is evidence that the protein: 327
carbohydrate ratio of the diet significantly determines lifetime egg production and life expectancy rather 328
than dietary restriction 99, 100. For instance, the egg production rate remains constant until the 329
carbohydrate content of the diet exceeds the protein content by 2-fold. Similarly, there is a continuous 330
decline in life span when the protein: carbohydrate ratio rises but a marked increase in life expectancy 331
if protein: carbohydrate ratio is low (about 1/10 or lower) 98, 99, 101, 102. Interestingly, similar effects of 332
macronutrient balance, so called “nutritional geometry”, on life span and reproductive capability are 333
reported in mice 103. Therefore, researchers are advised to carefully consider the aims and anticipated 334
outcomes of their dietary intervention studies in D. melanogaster and make use of the Geometric 335
Framework approach 102 so that they choose the most suitable food preparation for life span and 336
reproduction studies. 337
Biomarkers to be monitored in response to dietary factors 338
a. Life span 339
Drosophila has been used in genetic studies for decades but has only become interesting for nutrition-340
dependent intervention studies recently. The number of publications investigating the effects of dietary 341
factors on life span, physical fitness, fecundity and development has markedly increased within the last 342
ten years. Therefore, the fruit fly is an excellent model organism to study putative changes to life 343
expectancy caused by dietary factors. The fruit fly features easy husbandry, is relatively cheap, allows 344
for a high sample number for optimal power of experiments and exhibits a rapid reproduction and 345
lifecycle with a comparatively short lifetime. Furthermore, there are lots of mutant strains for the 346
15
mechanistic validation of initial findings. Similar to humans, where on average females live longer than 347
males 104, female flies often exhibit a prolonged life span compared with their male counterparts (Fig. 348
3A,B; 105). Life expectancy depends on genotype, female fecundity and mating status 106 and, 349
remarkably, sex-dependent differences in intestinal stem cell activity and background systemic 350
inflammation also determine longevity differences 107. 351
Moreover, male and female D. melanogaster from the same strain show diverging effects on life span 352
that are dependent on dietary interventions. Various isolated plant compounds, complex plant extracts 353
and macronutrients have recently been investigated for their life span-modulating abilities and have 354
revealed both beneficial and detrimental effects or have even been ineffective. Researchers have to take 355
into account that different fly strains exhibit variances in their genetically determined life expectancy 356
and in their susceptibility towards dietary interventions. Therefore, the use of various substance 357
concentrations and the inclusion of diverse fly strains in nutritional intervention studies are 358
recommended. 359
Importantly, both Drosophila adults and larva display learning processes and associate foods to their 360
nutritional values and tasting resulting in utilization or avoidance of a given food. As this ability can 361
cause confounding effects, especially in long-term feeding experiments like lifespan determination, the 362
performance of gustatory and choice assays (as described above) is highly recommended. Moreover, 363
larval feeding habits may affect the adult phenotype of D. melanogaster. It has been shown, that 364
administration of high contents of short carbohydrates, e.g., sucrose and fructose, can cause an obese 365
phenotype in mature fruit flies 108, 109. Researchers should consider the most suitable dietary composition 366
for their nutritional studies in respect of the desired phenotype and research question. 367
b. Expression of longevity-associated genes 368
The fruit fly possesses several longevity-associated genes that are related to nutrient dependent signaling 369
pathways and may be affected by caloric restriction. The expression of these longevity-associated genes 370
can further be affected by dietary secondary plant metabolites and nutrients resulting in life span 371
modification. Some promising candidate genes are spargel (srl), sirtuin 2 (sir2), chico, methuselah (mth) 372
and forkhead box, sub-group O (foxo). Thereby, both induction and down-regulation of the expression 373
16
of these genes may change the life span of the fruit fly. On the one hand, upregulation of srl and sir2 374
transcript levels extend life span while downregulation of chico and mth or solely targeted 375
overexpression of foxo at an adult age contributes to an extended life span. 376
Overexpression of srl, the Drosophila ortholog to mammalian PGC1, is associated with an increased 377
life span in Drosophila. The induction of srl transcription in intestinal stem and progenitor cells is 378
especially correlated with a significant extension in life span 110 while reduced srl expression causes a 379
decrease in life span 111. Similarly, sir2, a sirtuin 1 ortholog, is necessarily involved in life span 380
regulation in Drosophila as decreased expression of sir2 causes early lethality 112 while moderate 381
overexpression of sir2, especially in the fat body and neurons, prolongs the life span of fruit flies 82, 113. 382
Importantly, the mRNA expression of both srl and sir2 can be induced by dietary administration of plant 383
bioactives, e.g., the green tea catechin epigallocatechin gallate 78 and the isoflavone prunetin 20 that 384
contribute to life span extension in fruit flies. The Drosophila foxo, an ortholog of mammalian Foxo3a, 385
prolongs life span when its expression is exclusively activated in the pericerebral fat body resulting in a 386
reduction of insulin-like peptide 2 transcription in the neuronal insulin-producing cells 114. Similarly, life 387
span is elongated when foxo is overexpressed in the fly’s abdominal fat body 115, 116. Foxo overexpression 388
contributes to a reduction of the insulin/insulin growth factor-like signaling pathway, leading to life span 389
extension. Therefore, the induction of foxo during early adulthood in flies is most effective extending 390
life span compared to foxo upregulation at older ages 116. By contrast, the loss of chico expression, an 391
insulin receptor substrate in Drosophila mediating insulin/insulin-like growth factor signaling, leads to 392
longevity 117. This effect is putatively associated with the chico-dependent olfactory associative learning 393
ability and cognitive function of the fruit fly 118 and may be related to the fly’s nutritional behavior. 394
Likewise, the life span of Drosophila increases when flies carry a truncated version of mth 119 or when 395
mth expression is impaired by the use of specific mth inhibitors 120. The RNAi-mediated knockdown of 396
mth expression in insulin-producing cells in the fly’s brain is sufficient to extend its life span through 397
modification of the expression and release of insulin-like peptides and reduction of insulin/IGF signaling 398
121. As mth is assumed to act as a class B (secretin-like) G protein–coupled receptor involved in the drug 399
response and transduction of odorant stimuli, among other functions 119, the dietary modulation of mth 400
17
expression and function and the induction of longevity is conceivable. Notably, the effects of mth are 401
foxo-dependent and interactive, extending the life span of Drosophila 121. Besides the abovementioned 402
longevity-associated genes, many other genes may affect the life expectancy of Drosophila as 403
ascertained in mutant fly experiments 122. Tab. 1 gives a short overview on genes encoding proteins that 404
are putatively involved in life span regulation in the fruit fly. 405
c. Locomotor activity 406
Ingestion of plant bioactives may alter the metabolism and general health status of Drosophila as 407
reflected by changes in the movement behavior of the flies. Locomotor activity can be easily assessed 408
by performing the rapid iterative negative geotaxis (RING) assay. This method was introduced by 409
Gargano et al. 123 and takes advantage of the negative geotaxis of Drosophila. The assay is based on the 410
successive induction of climbing by tapping the flies onto the bottom of clear experimental vials (Fig. 411
3C) and letting them climb up the walls for a defined period of time. The higher the flies climb within 412
that interval the better their estimated overall health status, quantified by calculating the average 413
climbing score. The assay is useful for estimating the physical condition of fruit flies dependent on age, 414
genetic and dietary factors 20, 124-128. The advantages of this procedure are its sensitivity and the ability 415
to examine a large number of flies at the same time. 416
Another way to quantify spontaneous locomotor activity in Drosophila is the use of a PC-based 417
locomotor activity monitoring system. This method is based on using the interruption of an infrared light 418
beam gate by the fly to record the locomotor activity of individual flies and is costlier in terms of 419
equipment. However, this method possesses the advantage of being able to monitor the fly’s locomotor 420
behavior over longer periods of time under standardized housing conditions, and allows for the 421
examination of both circadian rhythm and sleep/rest parameters 129-131. 422
d. Intestinal barrier function 423
Gut barrier function is closely correlated with the overall life expectancy and health status of Drosophila 424
12, 123, 132, 133. The epithelial surface of the gut serves as a first-line of defense against microorganisms by 425
producing anti-microbial peptides (AMPs). Moreover, gut health and intestinal epithelial integrity 426
depend on appropriate stem cell proliferation and tissue homeostasis 134, which warrant gut integrity 110. 427
18
Adherence and septate junction proteins, e.g., armadillo, catenins, coracle, paxillin, polychaetoid, 428
shotgun and spectrins among others, are important for ensuring a strong gut barrier function in the fruit 429
fly that is comparable to that of mammals 135. Importantly, the loss of intestinal integrity causes a 430
shortening of the medium and maximum life spans of D. melanogaster 58, 136 while intestinal barrier 431
dysfunction increases with age 136 and predicts age-onset mortality 58. Premature mortality is also 432
associated with increased AMP expression 58 related to changes in the intestinal immune response. 433
Furthermore, alterations of the gut microbiome and development of midgut dysplasia that is reported for 434
some mutant fly strains are associated with increased gut dysfunction and barrier loss 137 and lead to a 435
shortened life span compared with wild-type flies 138. As dietary factors are able to affect the fly’s 436
metabolism, gene expression and microbiome composition 20, 78, 139-141, they may in turn modulate health 437
and life span in Drosophila via modification of gut integrity. 438
Visualization of gut integrity/leakage using food dyes 439
The Smurf assay is a non-invasive method of evaluating gut integrity in the fruit fly 58, 142 in response to 440
dietary factors. Therefore, flies are fed a specific diet for at least 30 days followed by concurrent dietary 441
administration of the food medium administered with a food dye, e.g., Brilliant Blue FCF (E133) or 442
Fluorescein (2.5% w/v) that normally do not traverse the gut for 9h to 7d 58, 110. Flies are visually 443
categorized into the smurf phenotype, by displaying intense body coloring indicating a leaky gut, or the 444
non-smurf phenotype with sole staining of the proboscis, anus and the gut itself, representing flies with 445
an intact intestinal barrier (Tab. 2, upper panel). Defined populations are scored according to these 446
phenotypes and the percentage change of the ratio of healthy flies to flies with impaired gut integrity is 447
calculated, highlighting the potential increase in health status by the dietary compound of interest. 448
Determination of intestinal immune function by quantification of anti-microbial peptide (AMP) 449
expression and stem cell proliferation 450
Whether plant bioactives and nutrient composition affect the health and life span of Drosophila, and to 451
what extent, potentially by their anti-bacterial or immuno-modulatory activities, can be evaluated by 452
various methods that give some indication of alterations in the immune status and function of the fly. 453
Therefore, flies are generally pre-fed with the experimental diet for a defined time and analyzed for AMP 454
19
expression, stem cell proliferation and bacterial load in the intestine. Unchallenged, compound-treated 455
and pathogen-infected flies can be generated for AMPs and stem cell proliferation targeted experiments. 456
Oral infection of fruit flies with Drosophila pathogens significantly induces AMP expression, e.g., 457
Metchnikowin, Diptericin, Attacins and Drosocin, in the gut that can be quantified by qRT-PCR 107, 135,
458
143-145. Accordingly, the repair of the epithelium is an important trait for survival following exposure to 459
intestinal stressors. Similar to mammals, the fruit fly possesses multipotent intestinal stem cells (ISC) in 460
the gut. ISC responsiveness to gut damage maintains intestinal homeostasis and promotes survival in 461
young flies while dysregulated intestinal stem ISC division may be detrimental at older ages 143, 146, 147. 462
Phospho-Histone H3 (PH3) staining of the gut is an appropriate method for depicting intestinal stem cell 463
proliferation 135, 143, 145, 148. Mild infections with a low infection dose of Drosophila pathogens, e.g., 464
Pectobacterium carotovorum subsp. carotovorum 15 and Pseudomonas entomophila L48, may induce 465
stem cell proliferation (Tab. 2, lower panel) while severe infections with high infection doses irreparably 466
damage the gut, resulting in distinctly decreased stem cell proliferation 143, 145. 467
Both gut-derived AMP expression and intestinal epithelial mitosis increase with age in the fruit fly 107 468
underlining the substantial importance of intestinal stem cell activity and background systemic 469
inflammation on the determination of Drosophila’s life expectancy 107. The treatment of flies with 470
secondary plant metabolites may affect the inflammatory response, AMP expression and stem cell 471
proliferation making these methods suitable tools for monitoring the effects of dietary supplements on 472
the health status of Drosophila. Notably, the indigenous gut microbiota activate a basal level of ISC 473
activity in the fruit fly 148, emphasizing the importance the gut microbiome for the immune and health 474
statuses of Drosophila. 475
Impact of gut microbiota and infection studies 476
The fly microbiome. The Drosophila gastrointestinal tract is colonized by various commensal 477
microorganisms similar to the mammalian gut. However, the fly gut contains a limited number of 478
microorganisms from approximately 30 species compared with more than 500 different bacterial species 479
in mammals; thus the fly gut displays a lower bacterial diversity 149, 150. Several human diseases including 480
inflammatory bowel disease, obesity and cardiovascular diseases have been connected to a detrimental 481
20
change in the gut microbiota 151. Therefore, model organisms hosting a rather simple microbiota are 482
needed to elucidate the underlying pathogenic mechanisms to establish preventive and therapeutic 483
strategies in the treatment of microbiota-associated diseases. 484
Most of the bacteria detected in the gut of fruit flies belong to the phyla Firmicutes and Proteobacteria. 485
Furthermore, Actinobacteria, Bacteroides and Cyanobacteria have been detected by 454 pyrosequencing 486
of 16S rRNA gene amplicons 152. The most dominating species of the Firmicutes and the Proteobacteria 487
belong to Lactobacillus and Acetobacter spp., respectively. The main species detected in the midguts of 488
flies are A. pomorum, A. tropicalis, L. frucitvorans, L. brevis and L. plantarum 152, 153. Importantly, the 489
composition of the microbiome depends on both the laboratory fly strain, sex and age of the flies 151. In 490
particular, frequently used strains differ in major bacterial species as Canton S flies hosted higher 491
numbers of Leuconostoc species whereas w1118 flies hosted higher numbers of Enterococcus species. 492
Furthermore, Lactobacillus species were more prominent in aged than in young flies and in Canton S 493
more than in w1118 flies. Age-dependent changes across fly strains and sex have also been demonstrated 494
for the bacterial species A. pasteurianus, L. plantarum and L. fructivorans 151. In yeast-deprived flies 495
being mono-associated with L. plantarum an increased larval growth as well as a reduction in 496
developmental timing has been documented being associated with an earlier metamorphosis 154. This L. 497
plantarum-induced accelerated growth is not detrimental to the flies as increased survival was detected 498
even in mono-associated male food-deprived flies 155. This study revealed that commensal L. plantarum 499
seems to be a beneficial member of the fly’s microbiome mediating an earlier hatching and an increased 500
life span without affecting fitness and reproduction. These results support the important role of the 501
intestinal microbiota in essential evolutionary processes. It has also been demonstrated that the 502
microbiota in Drosophila exert beneficial effects on the fly gut including the immune response, intestinal 503
physiology, gut function and gut homeostasis 137, 156-159. 504
Due to the presence of only a limited number of commensal aerotolerant bacteria in the gut, the fruit fly 505
offers an ideal research model to systematically unravel effects on the host-microbiota-interaction 151,
506
154, 160. As the fruit fly’s microbiota composition is mainly shaped by the host’s diet, the effect on the 507
microbiota of different compounds, nutrients and/or drugs can easily be investigated 161. Additionally, 508
21
the relatively cheap maintenance and the simple generation of germ-free and gnotobiotic animals makes 509
the fruit fly an important model organism to elucidate interactions between the host and its intestinal 510
microbiota 162. However, it should be taken into account that the microbiota of the fly depends on 511
permanent ingestion from the diet rather than on intestinal growth 153. Therefore, the total numbers of 512
flies kept per vial as well as the flipping procedure, should be carefully reconsidered for each experiment. 513
In addition to the environmental influence on the composition of microbiota in the fruit fly, the stomach-514
like copper cell region in the midgut of flies also contributes to the colonization as well as to the 515
composition of the intestinal microbiota 163. In addition to the mentioned properties, the fruit fly also 516
offers an ideal model organism to deeply investigate the effects on a potential host-microbiota-517
interaction as it combines both, genetic and experimental tractability. 518
Infection studies. The fruit fly is an appropriate model organism for investigating the preventive effects 519
of nutritional interventions on the severity of targeted infections, e.g., plant bioactives with putative anti-520
inflammatory or anti-bacterial activity. Several Drosophila pathogens induce either local or systemic 521
infections with mild or severe disease progression 164. Pathogens that induce mild infections in 522
Drosophila are Pectobacterium carotovorum subsp. carotovorum 15 165 and Serratia marcescens Db11 523
144 while Pseudomonas entomophila L48 induce severe disease 145, 166, 167. All of these pathogens can 524
either be introduced systemically or orally in high or low sub-lethal bacterial concentrations 145, 147, 167. 525
Simple experimental settings require the pre-feeding of flies for a defined time period before the flies 526
are orally or systemically exposed to pathogenic bacteria. Oral infections are then carried out by 527
application of a bacteria-sucrose-solution (OD600 =100 to 200) following starvation for 2h. For systemic 528
infections flies are infected by pricking into the thorax with a thin needle inoculated with a concentrated 529
bacterial pellet (4×1011 cells/ml at OD600 =200) 164, 165 or by microinjection of a defined volume of the 530
bacterial pellet using Nanoject™ and a pulled glass capillary 164. When performing nutritional 531
intervention studies targeting gut health in the fruit fly, researchers have to be aware that male 532
individuals display a higher level of systemic inflammation and are more susceptible to intestinal 533
infection than females, in general 107. Moreover, adequate bacterial uptake has to be ensured, which can 534
be ascertained by quantifying the bacterial load. Bacterial load assays are performed to estimate the 535
22
number of ingested pathogenic bacteria. Flies are starved for 2 h prior to the application of a bacterial-536
sucrose suspension, e.g., using GFP-labeled Erwinia carotovora carotovora 15 or Serratia marcescens 537
Db11 containing 0.5% FCF blue, for another 2 h. For quantification of the ingested amounts of bacteria, 538
flies are visually inspected under a fluorescent dissecting microscope when GFP-labeled bacteria are 539
used 164. Otherwise, flies are homogenized and their OD600 is measured spectrophotometrically. 540
Homogenized flies are plated onto lysogenic broth agar plates with selective antibiotics (if applicable) 541
and colony-forming units are counted following incubation at 29°C-30 °C overnight 107, 144. 542
Both wild-type and mutant fly strains, e.g., reporter flies carrying immune-inducible promoters of AMPs 543
ligated to green and red fluorescent proteins or LacZ-derived β-galactosidase activity 164, are suitable 544
infection models. Sensible and available read-outs are survivorship, AMPs expression, gut integrity, 545
stem cell proliferation and bacterial load/bacterial clearance. 546
e. Fertility 547
The sexual activity of the fruit fly affects its life expectancy and vice versa 168, 169. Fertility is a fitness 548
component and is correlated with life expectancy as germ cell ablation results in increased life 549
expectancy in D. melanogaster 170. Moreover, the expression of distinct hormone receptors affects life 550
span in a sex-specific manner 133. A reduction in fecundity is putatively associated with longevity in the 551
fruit fly, possibly mediated by dietary restriction-dependent effects 59, 171, 172. Interestingly, the nutritional 552
status of females affects their mating response 173, 174. Therefore, nutritional interventions e.g., 553
administration of secondary plant bioactives that may act as CR mimetics potentially affect the fertility 554
of female Drosophila. Male flies are able to transfer nutrients with their ejaculate 175 and thus may affect 555
female mating behavior with pheromone effective compounds 176. Testis development, accessory gland 556
protein synthesis and mating success are controlled by the Drosophila sex hormones 20-OH-ecdysone 557
and juvenile hormone 177, 178. Thus, plant bioactives exhibiting hormone-like structures might exhibit 558
estrogenic activity and affect fertility in the fruit fly. 559
Using D. melanogaster or C. elegans? – pros and cons 560
23
Invertebrate model organisms have become a cornerstone of various fields of biological and biomedical 561
research. Most prominent are the fruit fly D. melanogaster and the nematode Caenorhabditis elegans 562
that share many advantages listed in tab. 3 that permit systematic deciphering of the gene-gene and gene-563
environment interactions and often deliver fast answers to conserved fundamental biological problems. 564
Both, fly and worm are easy and cost-effective to grow in the laboratory. Their genome has been 565
sequenced and powerful forward and reverse genetics including RNAi and CRISPR are feasible. For 566
both models, a huge collection of mutants is available from community projects. Moreover, when 567
employing D. melanogaster or C. elegans usually one has to consider few ethical and regulatory 568
restrictions. Nevertheless, we have the opinion that the fruit fly provide some key features that favor it 569
as model with respect to nutritional research questions. (1) Despite its small body size, the fruit fly has 570
a relatively complex gastrointestinal system that, as in mammals, is highly specialized in anatomy, 571
organization and function. It consists of different cell types that enable e.g. frequent regeneration 572
processes. Compared to that, the gastrointestinal tract of C. elegans is very simple. It encompasses 573
twenty polyploid enterocytes and lacks enteroendocrine and stem cells. (2) As in mammals, the 574
alimentary tract of the fruit fly harbors various commensal microorganisms enabling the investigation 575
of host-microbiota-interaction. C. elegans lacks a microbiota. (3) The recent introduction of chemically 576
defined holidic media for D. melanogaster offers a broad range of opportunities to examine e.g. 577
micronutrient function and nutritional geometry 98,101. A comparable chemically defined medium for C. 578
elegans has not yet developed. (4) For nutritional studies it is often crucial to know how much food has 579
been ingested by an organism. However, appropriate methods for food uptake quantification have been 580
established only for D. melanogaster, but are rather lacking for C. elegans. (5) Furthermore, Drosophila 581
cell lines can be deemed as an extension of the fly model 179-181. Their culturing is easy and inexpensive 582
and a broad range of genetic, molecular and biochemical methodology has been established. Drosophila 583
cell cultures can be used to examine, e.g., cellular and biochemical response to environmental cues 584
including nutrients, bioactives and hormones. Up to now, no C. elegans cell culture lines exist. (6) 585
Finally, D. melanogaster contains a higher number of genes with orthologs in the human genome than 586
C. elegans. 587
24
Outlook – future directions 588
The fruit fly D. melanogaster is an emerging and valuable model organism for food and nutrition 589
research as there are numerous experimental methods, tools for analysis and genetic models available. 590
Nevertheless, specific aspects should be considered when designing future fruit fly nutrition studies to 591
improve experimental outcomes and study comparability. First, diet has a huge impact on many 592
biochemical, physiological and biological processes. A multitude of different, mainly undefined, fly 593
food recipes have been used to cultivate Drosophila. We encourage the Drosophila community to 594
develop standardized diets to ensure good reproducibility and comparability of the results between 595
nutritional intervention studies. Defined and purified experimental diets used for laboratory rodents 596
(including rats and mice) are already designed by the American Institute of Nutrition (AIN) 182, 183. In 597
particular, it may be useful to include a standardized defined holidic food medium. Recent progress in 598
the formulation of holidic fly media is very promising 92, 95. However, flies reared on these holidic food 599
preparations display a delay in larval development and a reduction of fecundity rates indicating that these 600
diets still have to be improved or adapted to the different life stages of D. melanogaster. The exact 601
nutrient requirements of the fruit fly still remain unsolved and must be elucidated in detail. The use of 602
holidic food preparations offers an excellent opportunity to prove the essentialness of micronutrients in 603
D. melanogaster development and physiology. Furthermore, the absorption of nutrients in the 604
gastrointestinal tract and the cellular uptake of metabolites, which represent key steps in alimentation, 605
as well as nutrient dependent transporters in the Drosophila gut are scarcely known and have to be 606
identified and characterized. Importantly, researchers have to take into account that the impact of 607
nutrients on biochemical, physiological and biological processes is affected by the genetic background 608
of an organism. As different fly strains show distinct inherent life expectancies and responsiveness to 609
dietary interventions, we recommend including at least two different wild-type fly strains in each key 610
experiment, where the effects of nutrients on D. melanogaster are examined. Additionally, it is essential 611
to ensure a sufficient replicate number and an adequate cohort size in certain experiments, such as life-612
span assays that are laborious, long and interference-prone, to create robust and solid data. Compared 613
with vertebrate models, D. melanogaster as a model organism offers a broad range of methodologies 614
25
(especially the versatile genetic toolbox) with which to approach basic problems of nutrition research in 615
a cost- and time-efficient way. The fruit fly occupies genes and proteins that are 50-60% orthologous to 616
mammalian ones 184-186 and that can be easily explored using FlyBase 187. Numerous mutant fly strains 617
are already commercially available at places such as the Bloomington Drosophila Stock Center (Indiana, 618
USA), the Drosophila Species Stock Center (California, USA), the Vienna Drosophila Resource Center 619
(Austria) and the Drosophila Genomics and Genetic Resources: Kyoto Stock Center (Japan). Further, 620
distinct target genes and signaling pathways can be specifically examined using RNAi and the 621
CRISPR/Cas system for genome engineering, as they can be used in the fruit fly. The fruit fly is an 622
appropriate model organism to target many aspects of basic nutritional research as central metabolic 623
pathways are evolutionarily well conserved. Moreover, D. melanogaster is an excellent model organism 624
to evaluate the impact of nutritional factors and diet composition on health and life span amongst many 625
other aspects in both wild type and mutant fly strains and in various disease models, e.g. fly strains 626
developing metabolic disorders like obesity, diabetes, immunological and motoric diseases188-191. 627
Results may be partly extrapolated to mammalian species. However, when the transfer of findings made 628
in the fruit fly to mammals is intended, verification in mammalian models, e.g., laboratory rodents, is 629
still required. 630
26
Abbreviations 631
foxo: forkhead box, sub-group O 632
IPC: insulin-producing cell 633
LCL: lower control limit 634
mth: methuselah 635
sir2: sirtuin 2 636
srl: spargel 637
UCL: upper control limit 638
Conflicts of interest 639
The authors declare no competing financial interest. 640
Funding sources 641
Source of funding: University of Kiel (federate state Schleswig-Holstein, Germany).
642
27
References 643
1. Chyb, S.; Gompel, N., Wild-type morphology. In Atlas of Drosophila Morphology: Wild-Type and 644
Classical Mutants, Academic Press: San Diego, 2013; pp 1-23. 645
2. Lemaitre, B.; Miguel-Aliaga, I., The digestive tract of Drosophila melanogaster. Annual review of 646
genetics 2013, 47, 377-404. 647
3. Prokop, A., Organs. droso4schools online resources for school lessons using the fruit fly 648
Drosophila. https://droso4schools.wordpress.com/organs/ (19.03.2018) 649
4. Gramates, L.; Marygold, S.; dos Santos, G.; Urbano, J.-M.; Antonazzo, G.; Matthews, B.; Rey, 650
A.; Tabone, C.; Crosby, M.; Emmert, D.; Falls, K.; Goodman, J.; Hu, Y.; Ponting, L.; Schroeder, 651
A.; Strelets, V.; Thurmond, J.; Zhou, P.; FlyBaseConsortium FlyBase: Organ System. 652
http://flybase.org/static_pages/imagebrowser/imagebrowser10.html (30.11.2017) 653
5. Castella, C.; Amichot, M.; Berge, J. B.; Pauron, D., DSC1 channels are expressed in both the 654
central and the peripheral nervous system of adult Drosophila melanogaster. Invertebrate 655
neuroscience : IN 2001, 4, 85-94. 656
6. Rein, K.; Zöckler, M.; Mader, M. T.; Grübel, C.; Heisenberg, M., The Drosophila Standard Brain. 657
Current Biology 2002, 12, 227-231. 658
7. Zheng, Z.; Lauritzen, J. S.; Perlman, E.; Robinson, C. G.; Nichols, M.; Milkie, D.; Torrens, O.; 659
Price, J.; Fisher, C. B.; Sharifi, N.; Calle-Schuler, S. A.; Kmecova, L.; Ali, I. J.; Karsh, B.; 660
Trautman, E. T.; Bogovic, J.; Hanslovsky, P.; Jefferis, G. S. X. E.; Kazhdan, M.; Khairy, K.; 661
Saalfeld, S.; Fetter, R. D.; Bock, D. D., A Complete Electron Microscopy Volume Of The Brain 662
Of Adult Drosophila melanogaster. bioRxiv 2017, 1-87. 663
https://www.biorxiv.org/content/early/2017/05/22/140905 (19.03.2018) 664
8. Shanbhag, S.; Tripathi, S., Epithelial ultrastructure and cellular mechanisms of acid and base 665
transport in the Drosophila midgut. J Exp Biol 2009, 212, 1731-44. 666
9. Strand, M.; Micchelli, C. A., Quiescent gastric stem cells maintain the adult Drosophila stomach. 667
Proceedings of the National Academy of Sciences of the United States of America 2011, 108, 668
17696-701. 669
10. Liu, Q.; Jin, L. H., Tissue-resident stem cell activity: a view from the adult Drosophila 670
gastrointestinal tract. Cell communication and signaling : CCS 2017, 15, 33. 671
11. Kuraishi, T.; Binggeli, O.; Opota, O.; Buchon, N.; Lemaitre, B., Genetic evidence for a protective 672
role of the peritrophic matrix against intestinal bacterial infection in Drosophila melanogaster. 673
Proceedings of the National Academy of Sciences of the United States of America 2011, 108, 674
15966-71. 675
12. Buchon, N.; Osman, D.; David, F. P.; Fang, H. Y.; Boquete, J. P.; Deplancke, B.; Lemaitre, B., 676
Morphological and molecular characterization of adult midgut compartmentalization in 677
Drosophila. Cell reports 2013, 3, 1725-38. 678
13. Harrop, T. W.; Pearce, S. L.; Daborn, P. J.; Batterham, P., Whole-genome expression analysis in 679
the third instar larval midgut of Drosophila melanogaster. G3 (Bethesda, Md.) 2014, 4, 2197-205. 680
14. Murakami, R.; Shigenaga, A.; Matsumoto, A.; Yamaoka, I.; Tanimura, T., Novel tissue units of 681
regional differentiation in the gut epithelium of Drosopbila, as revealed by P-element-mediated 682
detection of enhancer. Roux's archives of developmental biology : the official organ of the EDBO 683
1994, 203, 243-249. 684
15. Partridge, L.; Barrie, B.; Fowler, K.; French, V., Evolution and Development of Body Size and 685
Cell Size in Drosophila melanogaster in Response to Temperature. Evolution; international 686
journal of organic evolution 1994, 48, 1269-1276. 687
16. French, V.; Feast, M.; Partridge, L., Body size and cell size in Drosophila: the developmental 688
response to temperature. Journal of insect physiology 1998, 44, 1081-1089. 689
17. Turner, T. L.; Stewart, A. D.; Fields, A. T.; Rice, W. R.; Tarone, A. M., Population-Based 690
Resequencing of Experimentally Evolved Populations Reveals the Genetic Basis of Body Size 691
Variation in Drosophila melanogaster. PLOS Genetics 2011, 7, e1001336. 692
28
18. Shingleton, A. W.; Masandika, J. R.; Thorsen, L. S.; Zhu, Y.; Mirth, C. K., The sex-specific effects 693
of diet quality versus quantity on morphology in Drosophila melanogaster. Royal Society open 694
science 2017, 4, 170375. 695
19. Pitnick, S.; Garcia-Gonzalez, F., Harm to females increases with male body size in Drosophila 696
melanogaster. Proceedings. Biological sciences 2002, 269, 1821-8. 697
20. Piegholdt, S.; Rimbach, G.; Wagner, A. E., The phytoestrogen prunetin affects body composition 698
and improves fitness and lifespan in male Drosophila melanogaster. FASEB journal : official 699
publication of the Federation of American Societies for Experimental Biology 2016, 30, 948-58. 700
21. Tennessen, J. M.; Barry, W. E.; Cox, J.; Thummel, C. S., Methods for studying metabolism in 701
Drosophila. Methods (San Diego, Calif.) 2014, 68, 105-15. 702
22. Chen, Q.; Ma, E.; Behar, K. L.; Xu, T.; Haddad, G. G., Role of trehalose phosphate synthase in 703
anoxia tolerance and development in Drosophila melanogaster. The Journal of biological 704
chemistry 2002, 277, 3274-9. 705
23. Nghiem, D.; Gibbs, A. G.; Rose, M. R.; Bradley, T. J., Postponed aging and desiccation resistance 706
in Drosophila melanogaster. Experimental gerontology 2000, 35, 957-969. 707
24. Lamb, M. J., Age-related differences in the water content of normal and irradiated Drosophila 708
melanogaster. Experimental gerontology 1975, 10, 351-357. 709
25. Fairbanks, L. D.; Burch, G. E., Rate of water loss and water and fat content of adult Drosophila 710
melanogaster of different ages. Journal of insect physiology 1970, 16, 1429-36. 711
26. Chng, W. B.; Bou Sleiman, M. S.; Schupfer, F.; Lemaitre, B., Transforming growth factor 712
beta/activin signaling functions as a sugar-sensing feedback loop to regulate digestive enzyme 713
expression. Cell reports 2014, 9, 336-48. 714
27. Commin, C.; Aumont-Nicaise, M.; Claisse, G.; Feller, G.; Da Lage, J. L., Enzymatic 715
characterization of recombinant alpha-amylase in the Drosophila melanogaster species subgroup: 716
is there an effect of specialization on digestive enzyme? Genes & genetic systems 2013, 88, 251-717
9. 718
28. Gabrisko, M.; Janecek, S., Characterization of maltase clusters in the genus Drosophila. Journal 719
of molecular evolution 2011, 72, 104-18. 720
29. Davis, M. Y.; Trinh, K.; Thomas, R. E.; Yu, S.; Germanos, A. A.; Whitley, B. N.; Sardi, S. P.; 721
Montine, T. J.; Pallanck, L. J., Glucocerebrosidase Deficiency in Drosophila Results in alpha-722
Synuclein-Independent Protein Aggregation and Neurodegeneration. PLoS Genet 2016, 12, 723
e1005944. 724
30. Horne, I.; Haritos, V. S., Multiple tandem gene duplications in a neutral lipase gene cluster in 725
Drosophila. Gene 2008, 411, 27-37. 726
31. Sieber, M. H.; Thummel, C. S., The DHR96 nuclear receptor controls triacylglycerol homeostasis 727
in Drosophila. Cell Metab 2009, 10, 481-90. 728
32. Sieber, M. H.; Thummel, C. S., Coordination of triacylglycerol and cholesterol homeostasis by 729
DHR96 and the Drosophila LipA homolog magro. Cell Metab 2012, 15, 122-7. 730
33. Davis, C. A.; Riddell, D. C.; Higgins, M. J.; Holden, J. J.; White, B. N., A gene family in 731
Drosophila melanogaster coding for trypsin-like enzymes. Nucleic acids research 1985, 13, 6605-732
19. 733
34. Roman, G.; Meller, V.; Wu, K. H.; Davis, R. L., The opt1 gene of Drosophila melanogaster 734
encodes a proton-dependent dipeptide transporter. The American journal of physiology 1998, 275, 735
C857-69. 736
35. Palm, W.; Sampaio, J. L.; Brankatschk, M.; Carvalho, M.; Mahmoud, A.; Shevchenko, A.; Eaton, 737
S., Lipoproteins in Drosophila melanogaster--assembly, function, and influence on tissue lipid 738
composition. PLoS Genet 2012, 8, e1002828. 739
36. Southon, A.; Burke, R.; Camakaris, J., What can flies tell us about copper homeostasis? 740
Metallomics : integrated biometal science 2013, 5, 1346-56. 741
37. Xiao, G.; Zhou, B., What can flies tell us about zinc homeostasis? Archives of biochemistry and 742
biophysics 2016, 611, 134-141. 743
38. Quinn, L.; Yin, D.; Cranna, N.; Lee, A.; Mitchell, N.; Hannan, R., Steroid Hormones in 744
Drosophila: How Ecdysone Coordinates Developmental Signalling with Cell Growth and 745
Division. 2012. 746
29
39. Mirth, C. K.; Tang, H. Y.; Makohon-Moore, S. C.; Salhadar, S.; Gokhale, R. H.; Warner, R. D.; 747
Koyama, T.; Riddiford, L. M.; Shingleton, A. W., Juvenile hormone regulates body size and 748
perturbs insulin signaling in Drosophila. Proceedings of the National Academy of Sciences 2014, 749
111, 7018-7023. 750
40. Yamamoto, R.; Bai, H.; Dolezal, A. G.; Amdam, G.; Tatar, M., Juvenile hormone regulation of 751
Drosophila aging. BMC biology 2013, 11, 85. 752
41. Wu, Q.; Brown, M. R., Signaling and function of insulin-like peptides in insects. Annual review 753
of entomology 2006, 51, 1-24. 754
42. Nassel, D. R.; Vanden Broeck, J., Insulin/IGF signaling in Drosophila and other insects: factors 755
that regulate production, release and post-release action of the insulin-like peptides. Cellular and 756
molecular life sciences : CMLS 2016, 73, 271-90. 757
43. Psichas, A.; Reimann, F.; Gribble, F. M., Gut chemosensing mechanisms. The Journal of clinical 758
investigation 2015, 125, 908-17. 759
44. Chen, J.; Kim, S. M.; Kwon, J. Y., A Systematic Analysis of Drosophila Regulatory Peptide 760
Expression in Enteroendocrine Cells. Molecules and cells 2016, 39, 358-66. 761
45. Wegener, C.; Veenstra, J. A., Chemical identity, function and regulation of enteroendocrine 762
peptides in insects. Current opinion in insect science 2015, 11, 8-13. 763
46. Kim, S. K.; Rulifson, E. J., Conserved mechanisms of glucose sensing and regulation by 764
Drosophila corpora cardiaca cells. Nature 2004, 431, 316-20. 765
47. Song, W.; Cheng, D.; Hong, S.; Sappe, B.; Hu, Y.; Wei, N.; Zhu, C.; O'Connor, M. B.; Pissios, P.; 766
Perrimon, N., Midgut-Derived Activin Regulates Glucagon-like Action in the Fat Body and 767
Glycemic Control. Cell Metab 2017, 25, 386-399. 768
48. Nassel, D. R.; Liu, Y.; Luo, J., Insulin/IGF signaling and its regulation in Drosophila. General and 769
comparative endocrinology 2015, 221, 255-66. 770
49. Oldham, S., Obesity and nutrient sensing TOR pathway in flies and vertebrates: Functional 771
conservation of genetic mechanisms. Trends in endocrinology and metabolism: TEM 2011, 22, 772
45-52. 773
50. Maniere, G.; Ziegler, A. B.; Geillon, F.; Featherstone, D. E.; Grosjean, Y., Direct Sensing of 774
Nutrients via a LAT1-like Transporter in Drosophila Insulin-Producing Cells. Cell reports 2016, 775
17, 137-148. 776
51. Banerjee, K. K.; Deshpande, R. S.; Koppula, P.; Ayyub, C.; Kolthur-Seetharam, U., Central 777
metabolic sensing remotely controls nutrient-sensitive endocrine response in Drosophila via 778
Sir2/Sirt1-upd2-IIS axis. J Exp Biol 2017, 220, 1187-1191. 779
52. Delanoue, R.; Meschi, E.; Agrawal, N.; Mauri, A.; Tsatskis, Y.; McNeill, H.; Leopold, P., 780
Drosophila insulin release is triggered by adipose Stunted ligand to brain Methuselah receptor. 781
Science 2016, 353, 1553-1556. 782
53. Koyama, T.; Mirth, C. K., Growth-Blocking Peptides As Nutrition-Sensitive Signals for Insulin 783
Secretion and Body Size Regulation. PLoS biology 2016, 14, e1002392. 784
54. Sano, H.; Nakamura, A.; Texada, M. J.; Truman, J. W.; Ishimoto, H.; Kamikouchi, A.; Nibu, Y.; 785
Kume, K.; Ida, T.; Kojima, M., The Nutrient-Responsive Hormone CCHamide-2 Controls Growth 786
by Regulating Insulin-like Peptides in the Brain of Drosophila melanogaster. PLoS Genet 2015, 787
11, e1005209. 788
55. Diegelmann, S.; Jansen, A.; Jois, S.; Kastenholz, K.; Velo Escarcena, L.; Strudthoff, N.; Scholz, 789
H., The CApillary FEeder Assay Measures Food Intake in Drosophila melanogaster. 2017, 790
e55024. 791
56. Braco, J. T.; Gillespie, E. L.; Alberto, G. E.; Brenman, J. E.; Johnson, E. C., Energy-dependent 792
modulation of glucagon-like signaling in Drosophila via the AMP-activated protein kinase. 793
Genetics 2012, 192, 457-66. 794
57. Bross, T. G.; Rogina, B.; Helfand, S. L., Behavioral, physical, and demographic changes in 795
Drosophila populations through dietary restriction. Aging Cell 2005, 4, 309-17. 796
58. Rera, M.; Clark, R. I.; Walker, D. W., Intestinal barrier dysfunction links metabolic and 797
inflammatory markers of aging to death in Drosophila. Proceedings of the National Academy of 798
Sciences 2012, 109, 21528-21533. 799
30
59. Ja, W. W.; Carvalho, G. B.; Zid, B. M.; Mak, E. M.; Brummel, T.; Benzer, S., Water- and nutrient-800
dependent effects of dietary restriction on Drosophila lifespan. Proceedings of the National 801
Academy of Sciences of the United States of America 2009, 106, 18633-7. 802
60. Ghimire, S.; Kim, M. S., Enhanced Locomotor Activity Is Required to Exert Dietary Restriction-803
Dependent Increase of Stress Resistance in Drosophila. Oxidative medicine and cellular longevity 804
2015, 2015, 813801. 805
61. Bahadorani, S.; Bahadorani, P.; Phillips, J. P.; Hilliker, A. J., The effects of vitamin 806
supplementation on Drosophila life span under normoxia and under oxidative stress. The journals 807
of gerontology. Series A, Biological sciences and medical sciences 2008, 63, 35-42. 808
62. Deshpande, S. A.; Carvalho, G. B.; Amador, A.; Phillips, A. M.; Hoxha, S.; Lizotte, K. J.; Ja, W. 809
W., Quantifying Drosophila food intake: comparative analysis of current methodology. Nature 810
methods 2014, 11, 535-540. 811
63. Carvalho, G. B.; Kapahi, P.; Anderson, D. J.; Benzer, S., Allocrine modulation of feeding behavior 812
by the Sex Peptide of Drosophila. Current biology : CB 2006, 16, 692-6. 813
64. Carvalho, G. B.; Kapahi, P.; Benzer, S., Compensatory ingestion upon dietary restriction in 814
Drosophila melanogaster. Nat Methods 2005, 2, 813-5. 815
65. Ja, W. W.; Carvalho, G. B.; Mak, E. M.; de la Rosa, N. N.; Fang, A. Y.; Liong, J. C.; Brummel, 816
T.; Benzer, S., Prandiology of Drosophila and the CAFE assay. Proceedings of the National 817
Academy of Sciences of the United States of America 2007, 104, 8253-6. 818
66. Kimura, K.-I.; Shimozawa, T.; Tanimura, T., Isolation of Drosophila mutants with abnormal 819
proboscis extension reflex. Journal of Experimental Zoology 1986, 239, 393-399. 820
67. Toshima, N.; Tanimura, T., Taste preference for amino acids is dependent on internal nutritional 821
state in Drosophila melanogaster. The Journal of Experimental Biology 2012, 215, 2827-2832. 822
68. Wong, R.; Piper, M. D.; Wertheim, B.; Partridge, L., Quantification of food intake in Drosophila. 823
PLoS One 2009, 4, e6063. 824
69. Bezzar-Bendjazia, R.; Kilani-Morakchi, S.; Maroua, F.; Aribi, N., Azadirachtin induced larval 825
avoidance and antifeeding by disruption of food intake and digestive enzymes in Drosophila 826
melanogaster (Diptera: Drosophilidae). Pesticide biochemistry and physiology 2017, 143, 135-827
140. 828
70. Keita, S.; Masuzzo, A.; Royet, J.; Kurz, C. L., Drosophila larvae food intake cessation following 829
exposure to Erwinia contaminated media requires odor perception, Trpa1 channel and evf 830
virulence factor. Journal of insect physiology 2017, 99, 25-32. 831
71. Qi, W.; Yang, Z.; Lin, Z.; Park, J.-Y.; Suh, G. S. B.; Wang, L., A quantitative feeding assay in 832
adult Drosophila reveals rapid modulation of food ingestion by its nutritional value. Molecular 833
Brain 2015, 8, 87. 834
72. Peru y Colón de Portugal, R. L.; Ojelade, S. A.; Penninti, P. S.; Dove, R. J.; Nye, M. J.; Acevedo, 835
S. F.; Lopez, A.; Rodan, A. R.; Rothenfluh, A., Long-lasting, experience-dependent alcohol 836
preference in Drosophila. Addiction biology 2014, 19, 392-401. 837
73. El-Keredy, A.; Schleyer, M.; König, C.; Ekim, A.; Gerber, B., Behavioural Analyses of Quinine 838
Processing in Choice, Feeding and Learning of Larval Drosophila. PLOS ONE 2012, 7, e40525. 839
74. Kim, H.; Choi, M. S.; Kang, K.; Kwon, J. Y., Behavioral Analysis of Bitter Taste Perception in 840
Drosophila Larvae. Chemical senses 2016, 41, 85-94. 841
75. Mishra, D.; Miyamoto, T.; Rezenom, Y. H.; Broussard, A.; Yavuz, A.; Slone, J.; Russell, D. H.; 842
Amrein, H., The molecular basis of sugar sensing in Drosophila larvae. Current biology : CB 2013, 843
23, 1466-71. 844
76. Scherer, S.; Stocker, R. F.; Gerber, B., Olfactory learning in individually assayed Drosophila 845
larvae. Learning & memory (Cold Spring Harbor, N.Y.) 2003, 10, 217-25. 846
77. Lee, W. C.; Micchelli, C. A., Development and characterization of a chemically defined food for 847
Drosophila. PLoS One 2013, 8, e67308. 848
78. Wagner, A. E.; Piegholdt, S.; Rabe, D.; Baenas, N.; Schloesser, A.; Eggersdorfer, M.; Stocker, A.; 849
Rimbach, G., Epigallocatechin gallate affects glucose metabolism and increases fitness and 850
lifespan in Drosophila melanogaster. Oncotarget 2015, 6, 30568-30578. 851
31
79. Chandrashekara, K. T.; Shakarad, M. N., Aloe vera or Resveratrol Supplementation in Larval Diet 852
Delays Adult Aging in the Fruit Fly, Drosophila melanogaster. The Journals of Gerontology: 853
Series A 2011, 66A, 965-971. 854
80. Chandrashekara, K. T.; Popli, S.; Shakarad, M. N., Curcumin enhances parental reproductive 855
lifespan and progeny viability in Drosophila melanogaster. Age 2014, 36, 9702. 856
81. Lewis, E. B., A new standard food medium. Drosophila Information Service 1960, 34, 117-118. 857
82. Hoffmann, J.; Romey, R.; Fink, C.; Yong, L.; Roeder, T., Overexpression of Sir2 in the adult fat 858
body is sufficient to extend lifespan of male and female Drosophila. Aging (Albany NY) 2013, 5, 859
315-327. 860
83. Zhang, Z.; Han, S.; Wang, H.; Wang, T., Lutein extends the lifespan of Drosophila melanogaster. 861
Archives of gerontology and geriatrics 2014, 58, 153-9. 862
84. Wang, L.; Li, Y. M.; Lei, L.; Liu, Y.; Wang, X.; Ma, K. Y.; Zhang, C.; Zhu, H.; Zhao, Y.; Chen, 863
Z.-Y., Purple sweet potato anthocyanin attenuates fat-induced mortality in Drosophila 864
melanogaster. Experimental gerontology 2016, 82, 95-103. 865
85. Wood, J. G.; Rogina, B.; Lavu, S.; Howitz, K.; Helfand, S. L.; Tatar, M.; Sinclair, D., Sirtuin 866
activators mimic caloric restriction and delay ageing in metazoans. Nature 2004, 430, 686-689. 867
86. Si, H.; Fu, Z.; Babu, P. V.; Zhen, W.; Leroith, T.; Meaney, M. P.; Voelker, K. A.; Jia, Z.; Grange, 868
R. W.; Liu, D., Dietary epicatechin promotes survival of obese diabetic mice and Drosophila 869
melanogaster. The Journal of nutrition 2011, 141, 1095-100. 870
87. Baenas, N.; Piegholdt, S.; Schloesser, A.; Moreno, D. A.; García-Viguera, C.; Rimbach, G.; 871
Wagner, A. E., Metabolic Activity of Radish Sprouts Derived Isothiocyanates in Drosophila 872
melanogaster. International Journal of Molecular Sciences 2016, 17, 251. 873
88. Bauer, J. H.; Goupil, S.; Garber, G. B.; Helfand, S. L., An accelerated assay for the identification 874
of lifespan-extending interventions in Drosophila melanogaster. Proceedings of the National 875
Academy of Sciences of the United States of America 2004, 101, 12980-5. 876
89. Skorupa, D. A.; Dervisefendic, A.; Zwiener, J.; Pletcher, S. D., Dietary composition specifies 877
consumption, obesity and lifespan in Drosophila melanogaster. Aging cell 2008, 7, 478-490. 878
90. Grandison, R. C.; Piper, M. D.; Partridge, L., Amino-acid imbalance explains extension of lifespan 879
by dietary restriction in Drosophila. Nature 2009, 462, 1061-4. 880
91. Galenza, A.; Hutchinson, J.; Campbell, S. D.; Hazes, B.; Foley, E., Glucose modulates 881
&lt;em&gt;Drosophila&lt;/em&gt; longevity and immunity independent of the microbiota. 882
Biology Open 2016, 5, 165. 883
92. Piper, M. D. W.; Blanc, E.; Leitão-Gonçalves, R.; Yang, M.; He, X.; Linford, N. J.; Hoddinott, M. 884
P.; Hopfen, C.; Soultoukis, G. A.; Niemeyer, C.; Kerr, F.; Pletcher, S. D.; Ribeiro, C.; Partridge, 885
L., A holidic medium for Drosophila melanogaster. Nature methods 2014, 11, 886
10.1038/nmeth.2731. 887
93. Lee, W.-C.; Micchelli, C. A., Development and Characterization of a Chemically Defined Food 888
for Drosophila. PLoS ONE 2013, 8, e67308. 889
94. Sang, J. H., The Quantitative Nutritional Requirements of Drosophila Melanogaster. Journal of 890
Experimental Biology 1956, 33, 45-72. 891
95. Reis, T., Effects of Synthetic Diets Enriched in Specific Nutrients on Drosophila Development, 892
Body Fat, and Lifespan. PLoS ONE 2016, 11, e0146758. 893
96. Gospodaryov, D. V.; Yurkevych, I. S.; Jafari, M.; Lushchak, V. I.; Lushchak, O. V., Lifespan 894
extension and delay of age-related functional decline caused by Rhodiola roseadepends on dietary 895
macronutrient balance. Longevity & healthspan 2013, 2, 5. 896
97. Schriner, S. E.; Lee, K.; Truong, S.; Salvadora, K. T.; Maler, S.; Nam, A.; Lee, T.; Jafari, M., 897
Extension of Drosophila Lifespan by Rhodiola rosea through a Mechanism Independent from 898
Dietary Restriction. PLOS ONE 2013, 8, e63886. 899
98. Simpson, S. J.; Raubenheimer, D., Macronutrient balance and lifespan. Aging (Albany NY) 2009, 900
1, 875-80. 901
99. Lushchak, O. V.; Gospodaryov, D. V.; Rovenko, B. M.; Glovyak, A. D.; Yurkevych, I. S.; Klyuba, 902
V. P.; Shcherbij, M. V.; Lushchak, V. I., Balance Between Macronutrients Affects Life Span and 903
Functional Senescence in Fruit Fly Drosophila melanogaster. The Journals of Gerontology: Series 904
A 2012, 67A, 118-125. 905
32
100. Partridge, L.; Piper, M. D.; Mair, W., Dietary restriction in Drosophila. Mechanisms of ageing and 906
development 2005, 126, 938-50. 907
101. Lee, K. P.; Simpson, S. J.; Clissold, F. J.; Brooks, R.; Ballard, J. W. O.; Taylor, P. W.; Soran, N.; 908
Raubenheimer, D., Lifespan and reproduction in Drosophila: New insights from nutritional 909
geometry. Proceedings of the National Academy of Sciences 2008, 105, 2498. 910
102. Simpson, S. J.; Le Couteur, D. G.; Raubenheimer, D.; Solon-Biet, S. M.; Cooney, G. J.; Cogger, 911
V. C.; Fontana, L., Dietary protein, aging and nutritional geometry. Ageing research reviews 2017, 912
39, 78-86. 913
103. Solon-Biet, S. M.; Walters, K. A.; Simanainen, U. K.; McMahon, A. C.; Ruohonen, K.; Ballard, 914
J. W. O.; Raubenheimer, D.; Handelsman, D. J.; Le Couteur, D. G.; Simpson, S. J., Macronutrient 915
balance, reproductive function, and lifespan in aging mice. Proceedings of the National Academy 916
of Sciences 2015, 112, 3481. 917
104. Regan, J. C.; Partridge, L., Gender and longevity: Why do men die earlier than women? 918
Comparative and experimental evidence. Best Practice & Research Clinical Endocrinology & 919
Metabolism 2013, 27, 467-479. 920
105. Carazo, P.; Green, J.; Sepil, I.; Pizzari, T.; Wigby, S., Inbreeding removes sex differences in 921
lifespan in a population of Drosophila melanogaster. Biology Letters 2016, 12, 20160337. 922
106. Malick, L. E.; Kidwell, J. F., The Effect of Mating Status, Sex and Genotype on Longevity in 923
Drosophila melanogaster. Genetics 1966, 54, 203-209. 924
107. Regan, J. C.; Khericha, M.; Dobson, A. J.; Bolukbasi, E.; Rattanavirotkul, N.; Partridge, L., Sex 925
difference in pathology of the ageing gut mediates the greater response of female lifespan to dietary 926
restriction. eLife 2016, 5, e10956. 927
108. Rovenko, B. M.; Kubrak, O. I.; Gospodaryov, D. V.; Perkhulyn, N. V.; Yurkevych, I. S.; Sanz, 928
A.; Lushchak, O. V.; Lushchak, V. I., High sucrose consumption promotes obesity whereas its low 929
consumption induces oxidative stress in Drosophila melanogaster. Journal of insect physiology 930
2015, 79, 42-54. 931
109. Rovenko, B. M.; Perkhulyn, N. V.; Gospodaryov, D. V.; Sanz, A.; Lushchak, O. V.; Lushchak, V. 932
I., High consumption of fructose rather than glucose promotes a diet-induced obese phenotype in 933
Drosophila melanogaster. Comparative Biochemistry and Physiology Part A: Molecular & 934
Integrative Physiology 2015, 180, 75-85. 935
110. Rera, M.; Bahadorani, S.; Cho, J.; Koehler, C. L.; Ulgherait, M.; Hur, J. H.; Ansari, W. S.; Lo, T.; 936
Jones, D. L.; Walker, D. W., Modulation of longevity and tissue homeostasis by the Drosophila 937
PGC-1 homolog. Cell metabolism 2011, 14, 623-634. 938
111. Mukherjee, S.; Basar, M. A.; Davis, C.; Duttaroy, A., Emerging functional similarities and 939
divergences between Drosophila Spargel/dPGC-1 and mammalian PGC-1 protein. Frontiers in 940
genetics 2014, 5, 216. 941
112. Kusama, S.; Ueda, R.; Suda, T.; Nishihara, S.; Matsuura, E. T., Involvement of Drosophila Sir2-942
like genes in the regulation of life span. Genes & genetic systems 2006, 81, 341-8. 943
113. Rogina, B.; Helfand, S. L., Sir2 mediates longevity in the fly through a pathway related to calorie 944
restriction. Proceedings of the National Academy of Sciences of the United States of America 2004, 945
101, 15998-16003. 946
114. Hwangbo, D. S.; Gersham, B.; Tu, M.-P.; Palmer, M.; Tatar, M., Drosophila dFOXO controls 947
lifespan and regulates insulin signalling in brain and fat body. Nature 2004, 429, 562. 948
115. Giannakou, M. E.; Goss, M.; Junger, M. A.; Hafen, E.; Leevers, S. J.; Partridge, L., Long-lived 949
Drosophila with overexpressed dFOXO in adult fat body. Science 2004, 305, 361. 950
116. Giannakou, M. E.; Goss, M.; Jacobson, J.; Vinti, G.; Leevers, S. J.; Partridge, L., Dynamics of the 951
action of dFOXO on adult mortality in Drosophila. Aging Cell 2007, 6, 429-38. 952
117. Clancy, D. J.; Gems, D.; Harshman, L. G.; Oldham, S.; Stocker, H.; Hafen, E.; Leevers, S. J.; 953
Partridge, L., Extension of life-span by loss of CHICO, a Drosophila insulin receptor substrate 954
protein. Science 2001, 292, 104-6. 955
118. Naganos, S.; Horiuchi, J.; Saitoe, M., Mutations in the Drosophila insulin receptor substrate, 956
CHICO, impair olfactory associative learning. Neuroscience Research 2012, 73, 49-55. 957
119. Lin, Y.-J.; Seroude, L.; Benzer, S., Extended Life-Span and Stress Resistance in the Drosophila 958
Mutant methuselah. Science 1998, 282, 943-946. 959
33
120. Ja, W. W.; West Jr, A. P.; Delker, S. L.; Bjorkman, P. J.; Benzer, S.; Roberts, R. W., Extension of 960
Drosophila melanogaster life span with a GPCR peptide inhibitor. Nature Chemical Biology 2007, 961
3, 415. 962
121. Gimenez, L. E. D.; Ghildyal, P.; Fischer, K. E.; Hu, H.; Ja, W. W.; Eaton, B. A.; Wu, Y.; Austad, 963
S. N.; Ranjan, R., Modulation of Methuselah Expression Targeted to Drosophila Insulin-producing 964
Cells Extends Life and Enhances Oxidative Stress Resistance. Aging cell 2013, 12, 121-129. 965
122. Proshkina, E. N.; Shaposhnikov, M. V.; Sadritdinova, A. F.; Kudryavtseva, A. V.; Moskalev, A. 966
A., Basic mechanisms of longevity: A case study of Drosophila pro-longevity genes. Ageing 967
research reviews 2015, 24, 218-31. 968
123. Gargano, J. W.; Martin, I.; Bhandari, P.; Grotewiel, M. S., Rapid iterative negative geotaxis 969
(RING): a new method for assessing age-related locomotor decline in Drosophila. Experimental 970
gerontology 2005, 40, 386-395. 971
124. Wagner, A. E.; Piegholdt, S.; Rabe, D.; Baenas, N.; Schloesser, A.; Eggersdorfer, M.; Stocker, A.; 972
Rimbach, G., Epigallocatechin gallate affects glucose metabolism and increases fitness and 973
lifespan in Drosophila melanogaster. Oncotarget 2015, 6, 30568-78. 974
125. Ko, B. S.; Ahn, S. H.; Noh, D. O.; Hong, K.-B.; Han, S. H.; Suh, H. J., Effect of explosion-puffed 975
coffee on locomotor activity and behavioral patterns in Drosophila melanogaster. Food Research 976
International 2017, 100, 252-260. 977
126. Wang, H.-l.; Sun, Z.-o.; Rehman, R.-u.; Wang, H.; Wang, Y.-f.; Wang, H., Rosemary Extract-978
Mediated Lifespan Extension and Attenuated Oxidative Damage in Drosophila melanogaster Fed 979
on High-Fat Diet. Journal of Food Science 2017, 82, 1006-1011. 980
127. Siddique, Y. H.; Naz, F.; Jyoti, S.; Ali, F.; Fatima, A.; Rahul; Khanam, S., Protective effect of 981
Geraniol on the transgenic Drosophila model of Parkinson’s disease. Environmental Toxicology 982
and Pharmacology 2016, 43, 225-231. 983
128. Balasubramani, S. P.; Mohan, J.; Chatterjee, A.; Patnaik, E.; Kukkupuni, S. K.; Nongthomba, U.; 984
Venkatasubramanian, P., Pomegranate Juice Enhances Healthy Lifespan in Drosophila 985
melanogaster: An Exploratory Study. Frontiers in Public Health 2014, 2, 245. 986
129. Chiu, J. C.; Low, K. H.; Pike, D. H.; Yildirim, E.; Edery, I., Assaying Locomotor Activity to Study 987
Circadian Rhythms and Sleep Parameters in Drosophila. Journal of Visualized Experiments : JoVE 988
2010, 2157. 989
130. Pfeiffenberger, C.; Lear, B. C.; Keegan, K. P.; Allada, R., Locomotor activity level monitoring 990
using the Drosophila Activity Monitoring (DAM) System. Cold Spring Harbor protocols 2010, 991
2010, pdb.prot5518. 992
131. Woods, J. K.; Kowalski, S.; Rogina, B., Determination of the spontaneous locomotor activity in 993
Drosophila melanogaster. J Vis Exp 2014. 994
132. Simon, A. F.; Liang, D. T.; Krantz, D. E., Differential decline in behavioral performance of 995
Drosophila melanogaster with age. Mechanisms of ageing and development 2006, 127, 647-651. 996
133. Tricoire, H.; Battisti, V.; Trannoy, S.; Lasbleiz, C.; Pret, A. M.; Monnier, V., The steroid hormone 997
receptor EcR finely modulates Drosophila lifespan during adulthood in a sex-specific manner. 998
Mechanisms of ageing and development 2009, 130, 547-52. 999
134. Biteau, B.; Hochmuth, C. E.; Jasper, H., JNK activity in somatic stem cells causes loss of tissue 1000
homeostasis in the aging Drosophila gut. Cell stem cell 2008, 3, 442-55. 1001
135. Piegholdt, S.; Rimbach, G.; Wagner, A. E., Effects of the isoflavone prunetin on gut health and 1002
stress response in male Drosophila melanogaster. Redox Biology 2016, 8, 119-126. 1003
136. Ulgherait, M.; Rana, A.; Rera, M.; Graniel, J.; Walker, D. W., AMPK modulates tissue and 1004
organismal aging in a non-cell-autonomous manner. Cell reports 2014, 8, 1767-1780. 1005
137. Clark, R. I.; Salazar, A.; Yamada, R.; Fitz-Gibbon, S.; Morselli, M.; Alcaraz, J.; Rana, A.; Rera, 1006
M.; Pellegrini, M.; Ja, W. W.; Walker, D. W., Distinct Shifts in Microbiota Composition during 1007
Drosophila Aging Impair Intestinal Function and Drive Mortality. Cell reports 2015, 12, 1656-67. 1008
138. Guillou, A.; Troha, K.; Wang, H.; Franc, N. C.; Buchon, N., The Drosophila CD36 Homologue 1009
croquemort Is Required to Maintain Immune and Gut Homeostasis during Development and 1010
Aging. PLOS Pathogens 2016, 12, e1005961. 1011
34
139. Chandler, J. A.; Morgan Lang, J.; Bhatnagar, S.; Eisen, J. A.; Kopp, A., Bacterial Communities of 1012
Diverse Drosophila Species: Ecological Context of a Host–Microbe Model System. PLOS 1013
Genetics 2011, 7, e1002272. 1014
140. Sharon, G.; Segal, D.; Ringo, J. M.; Hefetz, A.; Zilber-Rosenberg, I.; Rosenberg, E., Commensal 1015
bacteria play a role in mating preference of Drosophila melanogaster. Proceedings of the National 1016
Academy of Sciences 2010, 107, 20051-20056. 1017
141. Wong, A. C.-N.; Wang, Q.-P.; Morimoto, J.; Senior, A. M.; Lihoreau, M.; Neely, G. G.; Simpson, 1018
S. J.; Ponton, F., Gut Microbiota Modifies Olfactory-Guided Microbial Preferences and Foraging 1019
Decisions in <em>Drosophila</em>. Current Biology 27, 2397-2404.e4. 1020
142. Dambroise, E.; Monnier, L.; Ruisheng, L.; Aguilaniu, H.; Joly, J. S.; Tricoire, H.; Rera, M., Two 1021
phases of aging separated by the Smurf transition as a public path to death. Scientific reports 2016, 1022
6, 23523. 1023
143. Buchon, N.; Broderick, N. A.; Poidevin, M.; Pradervand, S.; Lemaitre, B., Drosophila intestinal 1024
response to bacterial infection: activation of host defense and stem cell proliferation. Cell host & 1025
microbe 2009, 5, 200-11. 1026
144. Fink, C.; Hoffmann, J.; Knop, M.; Li, Y.; Isermann, K.; Roeder, T., Intestinal FoxO signaling is 1027
required to survive oral infection in Drosophila. Mucosal immunology 2016, 9, 927-36. 1028
145. Vallet-Gely, I.; Opota, O.; Boniface, A.; Novikov, A.; Lemaitre, B., A secondary metabolite acting 1029
as a signalling molecule controls Pseudomonas entomophila virulence. Cellular Microbiology 1030
2010, 12, 1666-1679. 1031
146. Biteau, B.; Karpac, J.; Supoyo, S.; Degennaro, M.; Lehmann, R.; Jasper, H., Lifespan extension 1032
by preserving proliferative homeostasis in Drosophila. PLoS Genet 2010, 6, e1001159. 1033
147. Chatterjee, M.; Ip, Y. T., Pathogenic Stimulation of Intestinal Stem Cell response in Drosophila. 1034
Journal of cellular physiology 2009, 220, 664-671. 1035
148. Buchon, N.; Broderick, N. A.; Chakrabarti, S.; Lemaitre, B., Invasive and indigenous microbiota 1036
impact intestinal stem cell activity through multiple pathways in Drosophila. Genes & 1037
Development 2009, 23, 2333-2344. 1038
149. Broderick, N. A.; Lemaitre, B., Gut-associated microbes of Drosophila melanogaster. Gut 1039
microbes 2012, 3, 307-21. 1040
150. Kuraishi, T.; Hori, A.; Kurata, S., Host-microbe interactions in the gut of Drosophila melanogaster. 1041
Frontiers in physiology 2013, 4, 375. 1042
151. Han, G.; Lee, H. J.; Jeong, S. E.; Jeon, C. O.; Hyun, S., Comparative Analysis of Drosophila 1043
melanogaster Gut Microbiota with Respect to Host Strain, Sex, and Age. Microbial ecology 2017, 1044
74, 207-216. 1045
152. Wong, C. N.; Ng, P.; Douglas, A. E., Low-diversity bacterial community in the gut of the fruitfly 1046
Drosophila melanogaster. Environmental microbiology 2011, 13, 1889-900. 1047
153. Blum, J. E.; Fischer, C. N.; Miles, J.; Handelsman, J., Frequent replenishment sustains the 1048
beneficial microbiome of Drosophila melanogaster. mBio 2013, 4, e00860-13. 1049
154. Storelli, G.; Defaye, A.; Erkosar, B.; Hols, P.; Royet, J.; Leulier, F., Lactobacillus plantarum 1050
promotes Drosophila systemic growth by modulating hormonal signals through TOR-dependent 1051
nutrient sensing. Cell Metab 2011, 14, 403-14. 1052
155. Tefit, M. A.; Leulier, F., Lactobacillus plantarum favors the early emergence of fit and fertile adult 1053
Drosophila upon chronic undernutrition. J Exp Biol 2017, 220, 900-907. 1054
156. Broderick, N. A.; Buchon, N.; Lemaitre, B., Microbiota-induced changes in drosophila 1055
melanogaster host gene expression and gut morphology. mBio 2014, 5, e01117-14. 1056
157. Erkosar, B.; Defaye, A.; Bozonnet, N.; Puthier, D.; Royet, J.; Leulier, F., Drosophila microbiota 1057
modulates host metabolic gene expression via IMD/NF-kappaB signaling. PLoS One 2014, 9, 1058
e94729. 1059
158. Erkosar, B.; Storelli, G.; Mitchell, M.; Bozonnet, L.; Bozonnet, N.; Leulier, F., Pathogen Virulence 1060
Impedes Mutualist-Mediated Enhancement of Host Juvenile Growth via Inhibition of Protein 1061
Digestion. Cell host & microbe 2015, 18, 445-55. 1062
159. Guo, L.; Karpac, J.; Tran, S. L.; Jasper, H., PGRP-SC2 promotes gut immune homeostasis to limit 1063
commensal dysbiosis and extend lifespan. Cell 2014, 156, 109-22. 1064
35
160. Ryu, J. H.; Kim, S. H.; Lee, H. Y.; Bai, J. Y.; Nam, Y. D.; Bae, J. W.; Lee, D. G.; Shin, S. C.; Ha, 1065
E. M.; Lee, W. J., Innate immune homeostasis by the homeobox gene caudal and commensal-gut 1066
mutualism in Drosophila. Science 2008, 319, 777-82. 1067
161. Leulier, F.; MacNeil, L. T.; Lee, W. J.; Rawls, J. F.; Cani, P. D.; Schwarzer, M.; Zhao, L.; Simpson, 1068
S. J., Integrative Physiology: At the Crossroads of Nutrition, Microbiota, Animal Physiology, and 1069
Human Health. Cell Metab 2017, 25, 522-534. 1070
162. Martino, M. E.; Ma, D.; Leulier, F., Microbial influence on Drosophila biology. Current opinion 1071
in microbiology 2017, 38, 165-170. 1072
163. Li, H.; Qi, Y.; Jasper, H., Preventing Age-Related Decline of Gut Compartmentalization Limits 1073
Microbiota Dysbiosis and Extends Lifespan. Cell host & microbe 2016, 19, 240-53. 1074
164. Neyen, C.; Bretscher, A. J.; Binggeli, O.; Lemaitre, B., Methods to study Drosophila immunity. 1075
Methods (San Diego, Calif.) 2014, 68, 116-28. 1076
165. Basset, A.; Khush, R. S.; Braun, A.; Gardan, L.; Boccard, F.; Hoffmann, J. A.; Lemaitre, B., The 1077
phytopathogenic bacteria Erwinia carotovora infects Drosophila and activates an immune 1078
response. Proceedings of the National Academy of Sciences of the United States of America 2000, 1079
97, 3376-81. 1080
166. Vodovar, N.; Vinals, M.; Liehl, P.; Basset, A.; Degrouard, J.; Spellman, P.; Boccard, F.; Lemaitre, 1081
B., Drosophila host defense after oral infection by an entomopathogenic Pseudomonas species. 1082
Proceedings of the National Academy of Sciences of the United States of America 2005, 102, 1083
11414-11419. 1084
167. Liehl, P.; Blight, M.; Vodovar, N.; Boccard, F.; Lemaitre, B., Prevalence of Local Immune 1085
Response against Oral Infection in a Drosophila/Pseudomonas Infection Model. PLOS Pathogens 1086
2006, 2, e56. 1087
168. Partridge, L.; Farquhar, M., Sexual activity reduces lifespan of male fruitflies. Nature 1981, 294, 1088
580. 1089
169. Partridge, L.; Green, A.; Fowler, K., Effects of egg-production and of exposure to males on female 1090
survival in Drosophila melanogaster. Journal of insect physiology 1987, 33, 745-749. 1091
170. Flatt, T.; Min, K. J.; D'Alterio, C.; Villa-Cuesta, E.; Cumbers, J.; Lehmann, R.; Jones, D. L.; Tatar, 1092
M., Drosophila germ-line modulation of insulin signaling and lifespan. Proceedings of the 1093
National Academy of Sciences of the United States of America 2008, 105, 6368-73. 1094
171. Bass, T. M.; Grandison, R. C.; Wong, R.; Martinez, P.; Partridge, L.; Piper, M. D., Optimization 1095
of dietary restriction protocols in Drosophila. The journals of gerontology. Series A, Biological 1096
sciences and medical sciences 2007, 62, 1071-81. 1097
172. Burger, J. M.; Buechel, S. D.; Kawecki, T. J., Dietary restriction affects lifespan but not cognitive 1098
aging in Drosophila melanogaster. Aging Cell 2010, 9, 327-35. 1099
173. Chapman, T.; Trevitt, S.; Partridge, L., Remating and male-derived nutrients in Drosophila 1100
melanogaster. Journal of evolutionary biology 1994, 7, 51-69. 1101
174. Fricke, C.; Bretman, A.; Chapman, T., Female nutritional status determines the magnitude and 1102
sign of responses to a male ejaculate signal in Drosophila melanogaster. Journal of evolutionary 1103
biology 2010, 23, 157-65. 1104
175. Markow, T. A.; Ankney, P. F., Drosophila males contribute to oogenesis in a multiple mating 1105
species. Science 1984, 224, 302-3. 1106
176. Averhoff, W. W.; Richardson, R. H., Multiple pheromone system controlling mating in Drosophila 1107
melanogaster. Proceedings of the National Academy of Sciences of the United States of America 1108
1976, 73, 591-3. 1109
177. Li, Y.; Ma, Q.; Cherry, C. M.; Matunis, E. L., Steroid signaling promotes stem cell maintenance 1110
in the Drosophila testis. Developmental biology 2014, 394, 129-41. 1111
178. Wolfner, M. F.; Partridge, L.; Lewin, S.; Kalb, J. M.; Chapman, T.; Herndon, L. A., Mating and 1112
hormonal triggers regulate accessory gland gene expression in male Drosophila. Journal of insect 1113
physiology 1997, 43, 1117-1123. 1114
179. Billmann, M.; Boutros, M., Methods for High-Throughput RNAi Screening in Drosophila Cells. 1115
Methods in molecular biology (Clifton, N.J.) 2016, 1478, 95-116. 1116
180. Moraes, A. M.; Jorge, S. A.; Astray, R. M.; Suazo, C. A.; Calderon Riquelme, C. E.; Augusto, E. 1117
F.; Tonso, A.; Pamboukian, M. M.; Piccoli, R. A.; Barral, M. F.; Pereira, C. A., Drosophila 1118
36
melanogaster S2 cells for expression of heterologous genes: From gene cloning to bioprocess 1119
development. Biotechnology advances 2012, 30, 613-28. 1120
181. Rodal, A. A.; Del Signore, S. J.; Martin, A. C., Drosophila comes of age as a model system for 1121
understanding the function of cytoskeletal proteins in cells, tissues, and organisms. Cytoskeleton 1122
(Hoboken, N.J.) 2015, 72, 207-24. 1123
182. Reeves, P. G.; Nielsen, F. H.; Fahey, G. C., Jr., AIN-93 purified diets for laboratory rodents: final 1124
report of the American Institute of Nutrition ad hoc writing committee on the reformulation of the 1125
AIN-76A rodent diet. The Journal of nutrition 1993, 123, 1939-51. 1126
183. Nutrition, A. I. o., Report of the American Institute of Nurtition ad hoc Committee on Standards 1127
for Nutritional Studies. The Journal of nutrition 1977, 107, 1340-8. 1128
184. Adams, M. D.; Celniker, S. E.; Holt, R. A.; Evans, C. A.; Gocayne, J. D.; Amanatides, P. G.; 1129
Scherer, S. E.; Li, P. W.; Hoskins, R. A.; Galle, R. F.; George, R. A.; Lewis, S. E.; Richards, S.; 1130
Ashburner, M.; Henderson, S. N.; Sutton, G. G.; Wortman, J. R.; Yandell, M. D.; Zhang, Q.; Chen, 1131
L. X.; Brandon, R. C.; Rogers, Y.-H. C.; Blazej, R. G.; Champe, M.; Pfeiffer, B. D.; Wan, K. H.; 1132
Doyle, C.; Baxter, E. G.; Helt, G.; Nelson, C. R.; Gabor, G. L.; Miklos; Abril, J. F.; Agbayani, A.; 1133
An, H.-J.; Andrews-Pfannkoch, C.; Baldwin, D.; Ballew, R. M.; Basu, A.; Baxendale, J.; 1134
Bayraktaroglu, L.; Beasley, E. M.; Beeson, K. Y.; Benos, P. V.; Berman, B. P.; Bhandari, D.; 1135
Bolshakov, S.; Borkova, D.; Botchan, M. R.; Bouck, J.; Brokstein, P.; Brottier, P.; Burtis, K. C.; 1136
Busam, D. A.; Butler, H.; Cadieu, E.; Center, A.; Chandra, I.; Cherry, J. M.; Cawley, S.; Dahlke, 1137
C.; Davenport, L. B.; Davies, P.; Pablos, B. d.; Delcher, A.; Deng, Z.; Mays, A. D.; Dew, I.; Dietz, 1138
S. M.; Dodson, K.; Doup, L. E.; Downes, M.; Dugan-Rocha, S.; Dunkov, B. C.; Dunn, P.; Durbin, 1139
K. J.; Evangelista, C. C.; Ferraz, C.; Ferriera, S.; Fleischmann, W.; Fosler, C.; Gabrielian, A. E.; 1140
Garg, N. S.; Gelbart, W. M.; Glasser, K.; Glodek, A.; Gong, F.; Gorrell, J. H.; Gu, Z.; Guan, P.; 1141
Harris, M.; Harris, N. L.; Harvey, D.; Heiman, T. J.; Hernandez, J. R.; Houck, J.; Hostin, D.; 1142
Houston, K. A.; Howland, T. J.; Wei, M.-H.; Ibegwam, C.; Jalali, M.; Kalush, F.; Karpen, G. H.; 1143
Ke, Z.; Kennison, J. A.; Ketchum, K. A.; Kimmel, B. E.; Kodira, C. D.; Kraft, C.; Kravitz, S.; 1144
Kulp, D.; Lai, Z.; Lasko, P.; Lei, Y.; Levitsky, A. A.; Li, J.; Li, Z.; Liang, Y.; Lin, X.; Liu, X.; 1145
Mattei, B.; McIntosh, T. C.; McLeod, M. P.; McPherson, D.; Merkulov, G.; Milshina, N. V.; 1146
Mobarry, C.; Morris, J.; Moshrefi, A.; Mount, S. M.; Moy, M.; Murphy, B.; Murphy, L.; Muzny, 1147
D. M.; Nelson, D. L.; Nelson, D. R.; Nelson, K. A.; Nixon, K.; Nusskern, D. R.; Pacleb, J. M.; 1148
Palazzolo, M.; Pittman, G. S.; Pan, S.; Pollard, J.; Puri, V.; Reese, M. G.; Reinert, K.; Remington, 1149
K.; Saunders, R. D. C.; Scheeler, F.; Shen, H.; Shue, B. C.; Sidén-Kiamos, I.; Simpson, M.; 1150
Skupski, M. P.; Smith, T.; Spier, E.; Spradling, A. C.; Stapleton, M.; Strong, R.; Sun, E.; Svirskas, 1151
R.; Tector, C.; Turner, R.; Venter, E.; Wang, A. H.; Wang, X.; Wang, Z.-Y.; Wassarman, D. A.; 1152
Weinstock, G. M.; Weissenbach, J.; Williams, S. M.; Woodage, T.; Worley, K. C.; Wu, D.; Yang, 1153
S.; Yao, Q. A.; Ye, J.; Yeh, R.-F.; Zaveri, J. S.; Zhan, M.; Zhang, G.; Zhao, Q.; Zheng, L.; Zheng, 1154
X. H.; Zhong, F. N.; Zhong, W.; Zhou, X.; Zhu, S.; Zhu, X.; Smith, H. O.; Gibbs, R. A.; Myers, 1155
E. W.; Rubin, G. M.; Venter, J. C., The Genome Sequence of Drosophila melanogaster. Science 1156
2000, 287, 2185. 1157
185. Reiter, L. T.; Potocki, L.; Chien, S.; Gribskov, M.; Bier, E., A Systematic Analysis of Human 1158
Disease-Associated Gene Sequences In Drosophila melanogaster. Genome Research 2001, 11, 1159
1114-1125. 1160
186. Fortini, M. E.; Skupski, M. P.; Boguski, M. S.; Hariharan, I. K., A Survey of Human Disease Gene 1161
Counterparts in the Drosophila Genome. The Journal of Cell Biology 2000, 150, 23-30. 1162
187. Gramates, L. S.; Marygold, S. J.; Santos, G. d.; Urbano, J.-M.; Antonazzo, G.; Matthews, B. B.; 1163
Rey, A. J.; Tabone, C. J.; Crosby, M. A.; Emmert, D. B.; Falls, K.; Goodman, J. L.; Hu, Y.; 1164
Ponting, L.; Schroeder, A. J.; Strelets, V. B.; Thurmond, J.; Zhou, P., FlyBase at 25: looking to 1165
the future. Nucleic acids research 2017, 45, D663-D671. 1166
188. de Paula, M. T.; Silva, M. R. P.; Araujo, S. M.; Bortolotto, V. C.; Martins, I. K.; Macedo, G. E.; 1167
Franco, J. L.; Posser, T.; Prigol, M., Drosophila melanogaster: A model to study obesity effects on 1168
genes expression and developmental changes on descendants. Journal of cellular biochemistry 1169
2018. 1170
189. Garschall, K.; Flatt, T., The interplay between immunity and aging in Drosophila. F1000Research 1171
2018, 7, 160. 1172
37
190. Hoffmann, J.; Romey, R.; Fink, C.; Roeder, T., Drosophila as a model to study metabolic disorders. 1173
Advances in biochemical engineering/biotechnology 2013, 135, 41-61. 1174
191. Musselman, L. P.; Kühnlein, R. P., Drosophila as a model to study obesity and metabolic disease. 1175
The Journal of Experimental Biology 2018, 221. 1176
192. Hartenstein, V., Atlas of Drosophila Development (Central Nervous System). The Interactive Fly 1177
1993, 10-11. 1178
193. Mair, W., Tipping the Energy Balance toward Longevity. Cell Metabolism 17, 5-6. 1179
194. Stenesen, D.; Suh, J. M.; Seo, J.; Yu, K.; Lee, K. S.; Kim, J. S.; Min, K. J.; Graff, J. M., Adenosine 1180
nucleotide biosynthesis and AMPK regulate adult life span and mediate the longevity benefit of 1181
caloric restriction in flies. Cell Metab 2013, 17, 101-12. 1182
195. Bai, H.; Kang, P.; Hernandez, A. M.; Tatar, M., Activin Signaling Targeted by Insulin/dFOXO 1183
Regulates Aging and Muscle Proteostasis in Drosophila. PLOS Genetics 2013, 9, e1003941. 1184
196. Simonsen, A.; Cumming, R. C.; Brech, A.; Isakson, P.; Schubert, D. R.; Finley, K. D., Promoting 1185
basal levels of autophagy in the nervous system enhances longevity and oxidant resistance in adult 1186
Drosophila. Autophagy 2008, 4, 176-84. 1187
197. Schriner, S. E.; Kuramada, S.; Lopez, T. E.; Truong, S.; Pham, A.; Jafari, M., Extension of 1188
Drosophila lifespan by cinnamon through a sex-specific dependence on the insulin receptor 1189
substrate chico. Experimental gerontology 2014, 60, 220-230. 1190
198. Tu, M. P.; Epstein, D.; Tatar, M., The demography of slow aging in male and female Drosophila 1191
mutant for the insulin-receptor substrate homologue chico. Aging Cell 2002, 1, 75-80. 1192
199. Yamamoto, R.; Tatar, M., Insulin receptor substrate chico acts with the transcription factor FOXO 1193
to extend Drosophila lifespan. Aging Cell 2011, 10, 729-32. 1194
200. Bai, H.; Kang, P.; Tatar, M., Drosophila insulin-like peptide-6 (dilp6) expression from fat body 1195
extends lifespan and represses secretion of Drosophila insulin-like peptide-2 from the brain. Aging 1196
Cell 2012, 11, 978-85. 1197
201. Giannakou, M. E.; Goss, M.; Partridge, L., Role of dFOXO in lifespan extension by dietary 1198
restriction in Drosophila melanogaster: not required, but its activity modulates the response. Aging 1199
Cell 2008, 7, 187-98. 1200
202. Rogina, B.; Helfand, S., Indy Mutations and Drosophila Longevity. Frontiers in genetics 2013, 4. 1201
203. Wang, P.-Y.; Neretti, N.; Whitaker, R.; Hosier, S.; Chang, C.; Lu, D.; Rogina, B.; Helfand, S. L., 1202
Long-lived Indy and calorie restriction interact to extend life span. Proceedings of the National 1203
Academy of Sciences 2009, 106, 9262. 1204
204. Wang, L.; Li, Y. M.; Lei, L.; Liu, Y.; Wang, X.; Ma, K. Y.; Chen, Z. Y., Cranberry anthocyanin 1205
extract prolongs lifespan of fruit flies. Experimental gerontology 2015, 69, 189-95. 1206
205. Kirby, K.; Hu, J.; Hilliker, A. J.; Phillips, J. P., RNA interference-mediated silencing of Sod2 in 1207
Drosophila leads to early adult-onset mortality and elevated endogenous oxidative stress. 1208
Proceedings of the National Academy of Sciences of the United States of America 2002, 99, 16162-1209
7. 1210
206. Phillips, P. J.; Parkes, L. T.; Hilliker, A. J., Targeted neuronal gene expression and longevity in 1211
Drosophila. 2001; Vol. 35, p 1157-64. 1212
207. Sun, Y.; Yolitz, J.; Alberico, T.; Sun, X.; Zou, S., Lifespan Extension by Cranberry 1213
Supplementation Partially Requires SOD2 and is Life Stage Independent. Experimental 1214
gerontology 2014, 50, 57-63. 1215
208. Huang, C. W.; Wang, H. D.; Bai, H.; Wu, M. S.; Yen, J. H.; Tatar, M.; Fu, T. F.; Wang, P. Y., 1216
Tequila Regulates Insulin-Like Signaling and Extends Life Span in Drosophila melanogaster. The 1217
journals of gerontology. Series A, Biological sciences and medical sciences 2015, 70, 1461-9. 1218
209. Corsi, A. K.; Wightman, B.; Chalfie, M., A Transparent Window into Biology: A Primer on 1219
Caenorhabditis elegans. Genetics 2015, 200, 387-407. 1220
1221
38
Figure captions
Figure 1. (A) Anatomy of Drosophila melanogaster with a focus on the digestive tract (schematic).
Graph made according to Prokop and Hartenstein 3, 192. (B) Body weight and composition of the w1118
D. melanogaster wild type strain.
IPCs: insulin-producing cells; crop: represents the stomach; midgut: small intestine; hindgut: large
intestine and rectum; Malpighian tubules: kidney
Figure 2. Chemical structure of sulforhodamine B (left panel) and white light pictures of male and
female w1118 D. melanogaster that ingested sulforhodamine B with the fly food medium (0.2% w/v) for
500min (right panel) leading to an accumulation of the food dye in the fly’s fat body. Own unpublished
work.
Figure 3. Life span and locomotor activity in w1118 D. melanogaster reared on a standard sugar-yeast-
corn meal diet dependent on sex. (A) Representative life span curves of male and female w1118 D.
melanogaster that received a life-long standard diet. (B) Females had a longer mean, median and
maximum1 life span (longest-lived 10%) than males. (C) The rapid iterative negative geotaxis (RING)
assay is a common method used for the evaluation of the locomotor activity of D. melanogaster
dependent on age and nutritional interventions. Flies are repeatedly tapped to the bottom of clear vials
and allowed to climb up for a defined time period before a photograph is taken. Climbing activity is
quantified by scoring the distance that is overcome by the flies. The higher the climbing distance the
better the overall health status of the flies.
1maximum refers to the longest-lived 10%. Median life span -lower control limit: 42 days for males and
58 days for females; median life span -upper control limit: 47 days for males and 63 days for females.
Own unpublished work.
39
Figures
Figure 1
Figure 1. (A) Anatomy of Drosophila melanogaster with a focus on the digestive tract (schematic).
Graph made according to Prokop and Hartenstein 3, 192. (B) Body weight and composition of the w1118
D. melanogaster wild type strain.
IPCs: insulin-producing cells; crop: represents the stomach; midgut: small intestine; hindgut: large
intestine and rectum; Malpighian tubules: kidney
40
Figure 2
Figure 2. Chemical structure of sulforhodamine B (left panel) and white light pictures of male and
female w1118 D. melanogaster that ingested sulforhodamine B with the fly food medium (0.2% w/v) for
500min (right panel) leading to an accumulation of the food dye in the fly’s fat body. Own unpublished
work.
41
Figure 3
Figure 3. Life span and locomotor activity in w1118 D. melanogaster reared on a standard sugar-yeast-
corn meal diet dependent on sex. (A) Representative life span curves of male and female w1118 D.
melanogaster that received a life-long standard diet. (B) Females had a longer mean, median and
maximum1 life span (longest-lived 10%) than males. (C) The rapid iterative negative geotaxis (RING)
assay is a common method used for the evaluation of the locomotor activity of D. melanogaster
dependent on age and nutritional interventions. Flies are repeatedly tapped to the bottom of clear vials
and allowed to climb up for a defined time period before a photograph is taken. Climbing activity is
42
quantified by scoring the distance that is overcome by the flies. The higher the climbing distance the
better the overall health status of the flies.
1maximum refers to the longest-lived 10%. Median life span -lower control limit: 42 days for males and
58 days for females; median life span -upper control limit: 47 days for males and 63 days for females.
Own unpublished work.
43
Tables
Table 1
Table 1: Longevity-associated genes of Drosophila melanogaster whose expression may is controlled
by dietary factors.
Gene Full name Biological function Life-extending if Reference
AMPK
AMP-activated
protein kinase α
subunit
starvation response, lipid
metabolism, regulation of
digestive system processes, TOR
signaling
up-regulation/
activation 193, 194
Atg8a Autophagy-
related 8a
regulation of autophagy, life span
determination up-regulation 195, 196
Chico chico
regulation of immune response,
organ/organism growth,
metabolic processes, ageing
down-regulation 197-199
dILP6 Insulin-like
peptide 6 growth regulation overexpression 200
Foxo forkhead box,
sub-group O
carbohydrate metabolism,
circadian rhythm, regulation of
proliferation and growth,
hormone synthesis
overexpression 114, 115, 201
Indy I'm not dead yet triglyceride metabolism, citrate
cycle down-regulation 202, 203
Mth methuselah stress response, ageing
down-regulation/
partial loss-of-
function
119, 204
Sir2 Sirtuin 1 regulation of protein/histone
acetylation overexpression 20, 82
Sod2
Superoxide
dismutase 2
(Mn)
regulation of metabolic
processes, autophagy and life
span, hemocyte proliferation,
heart morphogenesis
up-regulation/
activation 83, 205-207
Srl spargel energy homeostasis overexpression 110
Teq Tequila glucose homeostatis, memory
function down-regulation 208
44
Table 2
Table 2. Intestinal integrity and stem cell proliferation can be visualized using the smurf assay and
phospho-histone H3 (PH3) immunofluorescence staining to indicate stem cell proliferation.
Strain/Sex/Age
Smurf assay Non-smurf smurf phenotype Dye used for
visualization
w1118
30 days
Brilliant Blue FCF
(E133)
Stem cell
proliferation Healthy (non-infected control)
Infected with Pseudomonas
entomophila at a low infectious
dose of OD600=5
w1118
10 days
Smurf assay: The flies were fed an experimental diet for 30 days prior to the dietary administration of an E133
Brilliant Blue FCF-supplemented diet (2.5% w/v) for 7 days. Subsequently, flies were anaesthetized with carbon
dioxide and the number of flies with or without the Smurf phenotype were counted.
Stem cell proliferation: The flies were fed an experimental diet for 10 days. The midguts were dissected and fixed
using a 4% paraformaldehyde solution (pH=7.4). Samples were incubated with an anti-phospho-histone H3
antibody (PH3) prior to incubation with an AlexaFluor594-conjugated secondary antibody and counterstaining
with DAPI. Images were acquired with a fluorescence microscope using TexasRed and DAPI filter systems as
previously described 135.
Pictures: Own unpublished work.
45
Table 3
Table 3: Comparison of the invertebrate models Drosophila melanogaster and Caenorhabditis elegans.
D. melanogaster C. elegans
Phylum Arthropoda Nematoda
Sexes females and males hermaphrodites and males
Chromosomes 3 pairs of autosomes (2-4)
+ an X/Y pair
5 pairs of autosomes (I-V)
+ 1 or 2 sex chromosomes
Genome size & gene
number 165 Mb, approx.. 17.000 100 Mb, approx.. 20,000
Orthologs in human
genome 50% 38%
Lifetime fecundity approx. 400 progeny
- approx. 300 self-progeny;
- up to 1000 progeny after
mating
Egg to adult
development
egg, instar larvae 1-3, pupa, adult
10 d at 25 °C
egg, L1 to L4 larvae, adult
3 d at 20 °C
Average lifespan approx. 50-60 d at 25 °C approx. 15-20 d at 20 °C
Laboratory diet
complex diets
defined holidic medium
E. coli lawn on agar plates
E. coli in complex liquid media
Axenic semi-defined medium
Anatomy of intestinal
tract
mouth, pharynx, esophagus, crop,
cardia, midgut, hindgut, anus mouth, pharynx, intestine, rectum
Intestinal cell-types
enterocytes, enteroendocrine
cells
Stem cells
20 polyploid epithelial cells
Microbiota
species of the phyla Firmicutes,
Proteobacteria, Actinobacteria,
Bacteroides and Cyanobacteria
lacking
Cell culture Available Not available
164, 191, 209
46
TOC Graphic
--- for table of contents only ---
47
Table of contents
ABSTRACT ............................................................................................................................................. 2
INTRODUCTION ................................................................................................................................... 3
FLY ANATOMY..................................................................................................................................... 3
A. MORPHOLOGY AND GUT ANATOMY ................................................................................................ 3
B. SIZE, BODY WEIGHT, BODY COMPOSITION ...................................................................................... 5
C. ENZYMATIC MACHINERY................................................................................................................ 6
D. NUTRIENT SENSING AND ENDOCRINE SIGNALING ........................................................................... 7
MEASUREMENT OF FEED INTAKE AND COMPOSITION OF EXPERIMENTAL DIETS 10
A. QUANTIFICATION OF FOOD INGESTION AND EXAMINATION OF MARKED FEED PREFERENCES ........ 10
B. COMPLEX VERSUS CHEMICALLY-DEFINED DIETS .......................................................................... 12
BIOMARKERS TO BE MONITORED IN RESPONSE TO DIETARY FACTORS ................... 14
A. LIFE SPAN ..................................................................................................................................... 14
B. EXPRESSION OF LONGEVITY-ASSOCIATED GENES ......................................................................... 15
C. LOCOMOTOR ACTIVITY ................................................................................................................ 17
D. INTESTINAL BARRIER FUNCTION ................................................................................................... 17
E. FERTILITY .................................................................................................................................... 22
USING D. MELANOGASTER OR C. ELEGANS? – PROS AND CONS ........................................ 22
OUTLOOK – FUTURE DIRECTIONS ............................................................................................. 24
REFERENCES ...................................................................................................................................... 27
FIGURE CAPTIONS ........................................................................................................................... 38
FIGURES ............................................................................................................................................... 39
TABLES ................................................................................................................................................. 43
TOC GRAPHIC .................................................................................................................................... 46
... The fruit fly D. melanogaster has proven to be a valuable model organism in nutrition research in recent decades (Staats et al., 2018;FIGURE 11 Schematic overview of the Gluc-HET experiment. (A) The location of the air bladder of a fertilized hen's egg (11 days old) is marked and test substancecontaining solution is applied into the air compartment. ...
... Compared to rodents, this invertebrate model organism has numerous advantages. Due to an affordable maintenance, a high reproduction rate and a relatively short lifespan of D. melanogaster, high-throughput analyses are possible (Staats et al., 2018;Jans et al., 2021). Moreover, as the hen's egg test, experiments can be carried out without serious ethical concerns and regulatory frameworks. ...
Article
Full-text available
Plant extracts are increasingly recognized for their potential in modulating (postprandial) blood glucose levels. In this context, root extracts are of particular interest due to their high concentrations and often unique spectrum of plant bioactives. To identify new plant species with potential glucose-lowering activity, simple and robust methodologies are often required. For this narrative review, literature was sourced from scientific databases (primarily PubMed) in the period from June 2022 to January 2024. The regulatory targets of glucose homeostasis that could be modulated by bioactive plant compounds were used as search terms, either alone or in combination with the keyword "root extract". As a result, we present a comprehensive methodological toolbox for studying the glucose homeostasis modulating properties of plant extracts and its constituents. The described assays encompass in-vitro investigations involving enzyme inhibition (α-amylase, α-glucosidase, dipeptidyl peptidase 4), assessment of sodium-dependent glucose transporter 1 activity, and evaluation of glucose transporter 4 translocation. Furthermore, we describe a patch-clamp technique to assess the impact of extracts on K ATP channels. While validating in-vitro findings in living organisms is imperative, we introduce two screenable in-vivo models (the hen's egg test and Drosophila melanogaster). Given that evaluation of the bioactivity of plant extracts in rodents and humans represents the current gold standard, we include approaches addressing this aspect.
... Drosophila melanogaster is a useful model to detect the relationship between diet and diseases. 23 Thus, Drosophila melanogaster was used to investigate the function and regulatory mechanism of COS on the intestine disruption induced by sodium dodecyl sulfate (SDS) 24,25 in vivo. The survival rate, intestinal homeostasis, and antioxidant capability were analyzed in female flies with the presence or absence of COS after SDS challenge. ...
... The fruit fly has been proved as an excellent in vivo model for dissecting mechanism of human diseases and screening of potential drugs. 23 It is a well-established genetic model organism and has been widely applied to explore almost all biological processes. 51 Moreover, it is more attractive for exploring mechanisms relevant to intestine as the simpler gut structure and lower diversity of gut microbiota in fruit flies than mammals. ...
Article
Full-text available
Recent evidence suggests the anti‐inflammatory effect of carrageenan oligosaccharides (COS). The effects of COS on intestinal injury induced by 0.6% sodium dodecyl sulfate (SDS) and the molecular mechanisms involved were investigated in this study. 0.625, 1.25, and 2.5 mg/mL COS in diet had no toxic effect in flies, and they could all prolong SDS‐treated female flies' survival rate. 1.25 mg/mL COS prevented the development of inflammation by improving the intestinal barrier integrity and maintaining the intestinal morphology stability, inhibited the proliferation of intestine stem cells (ISCs), and the production of lysosomes induced by SDS, accompanied by a decrease in the expression of autophagy‐related genes. Moreover, COS decreased the active oxygen species (ROS) content in gut and increased the antioxidant activity in SDS‐induced female flies, while COS still played a role in increasing survival rate and decreasing intestinal leakage in CncC‐RNAi flies. The improvement of anti‐inflammation capacity may be associated with the regulation of intestinal microflora with COS supplementation for Drosophila melanogaster. COS changed the gut microbiota composition, and COS had no effect on germ‐free (GF) flies. It is highlighted that COS could not work in Relish‐RNAi flies, indicating relish is required for COS to perform beneficial effects. These results provide insights into the study of gut microbiota interacting with COS to modulate intestinal inflammation in specific hosts.
... D. melanogaster is especially known for its use in genetic research but has also been well established as a model to investigate and answer physiological questions (18). Due to its short generation time, the relatively inexpensive and easy maintenance, and the fact that 60% of the genes present have orthologs in mammals, the fruit fly is an ideal model for studying evolutionarily conserved genes and signaling pathways (19,20). Due to the absence of an adaptive immune response in the fruit fly, processes of the innate immunity can be studied individually (21). ...
Article
Full-text available
Since infections with antibiotic-resistant bacteria cause increasing problems worldwide, the identification of alternative therapies is of great importance. Plant-derived bioactives, including allyl-isothiocyanate (AITC), have received attention for their antimicrobial properties. The present study therefore investigates the impact of AITC on survival and antimicrobial peptide (AMP) levels in Drosophila melanogaster challenged with the fly pathogenic bacteria Pectobacterium carotovorum subsp. carotovorum and Leuconostoc pseudomesenteroides. AITC, a sulfur-containing compound derived from glucosinolates, exhibits antimicrobial properties and has been suggested to modulate AMP expression. By using D. melanogaster, we demonstrate that AITC treatment resulted in a concentration-dependent decrease of survival rates among female flies, particularly in the presence of the Gram-negative bacterium Pectobacterium carotovorum subsp. carotovorum, whereas AITC did not affect survival in male flies. Despite the ability of isothiocyanates to induce AMP expression in cell culture, we did not detect significant changes in AMP mRNA levels in infected flies exposed to AITC. Our findings suggest sex-specific differences in response to AITC treatment and bacterial infections, underlining the complexity of host–pathogen interactions and potential limitations of AITC as a preventive or therapeutic compound at least in D. melanogaster models of bacterial infections.
Article
Drosophila melanogaster, or the fruit fly, is widely used for modeling numerous human diseases, such as neurodegeneration, tumor development, cachexia, and intestinal dysfunction. It is a suitable model organism for research targeting the physiology and pathophysiology of the intestinal epithelial barrier and has also been used as a model organism for preliminary drug and bioactive nutrient screening. However, the application of D. melanogaster in research on drug bioavailability and pharmacokinetic properties has not yet been well explored. In this study, we applied D. melanogaster to investigate the absorption and excretion of the orally administered phytoestrogens daidzein, glycitein, genistein, and their glycosides. Therefore, we established a quick, noninvasive method to quantify compound retention in D. melanogaster, suitable for the investigation of a broad variety of potentially bioactive substances. We showed that fruit fly sex plays a key role in the metabolization, transportation, and excretion of phytoestrogenic isoflavones. In particular, female fruit flies retained significantly more isoflavones than male fruit flies, which was reflected in the greater metabolic impact of isoflavones on females. Male fruit flies excreted more isoflavones than females did, which was linked to the upregulation of the xenobiotic transporter gene Mdr50. We also demonstrated that micellized isoflavones were more bioavailable than powdered isoflavones, independent of sex, age or the addition of dietary fibers.
Article
Full-text available
Lifespan is a complex quantitative trait involving genetic and non-genetic factors as well as the peculiarities of ontogenesis. As with all quantitative traits, lifespan shows considerable variation within populations and between individuals. Drosophila, a favourite object of geneticists, has greatly advanced our understanding of how different forms of variability affect lifespan. This review considers the role of heritable genetic variability, phenotypic plasticity and stochastic variability in controlling lifespan in Drosophila melanogaster. We discuss the major historical milestones in the development of the genetic approach to study lifespan, the breeding of long-lived lines, advances in lifespan QTL mapping, the environmental factors that have the greatest influence on lifespan in laboratory maintained flies, and the mechanisms, by which individual development affects longevity. The interplay between approaches to study ageing and lifespan limitation will also be discussed. Particular attention will be paid to the interaction of different types of variability in the control of lifespan.
Article
Ginseng is an ancient medicinal and edible plant with many health benefits, and can serve as a drug and dietary supplement, but there are few relevant studies on its use to ease ultraviolet (UV) irradiation damage. After 0.8 mg/mL ginseng extract (GE) was added to the medium of female Drosophila melanogaster subjected to UV irradiation, the lifespan, climbing ability, sex ratio, developmental cycle, and antioxidant capacity of flies were examined to evaluate the GE function. In addition, the underlying mechanism by which GE enhances the irradiation tolerance of D. melanogaster was explored. With GE supplementation, female flies subjected to UV irradiation exhibited an extension in their lifespan, enhancement in their climbing ability, improvement in their offspring sex ratio, and restoration of the normal development cycle by increasing their antioxidant activity. Finally, further experiments indicated that GE could enhance the irradiation tolerance of female D. melanogaster by upregulating the gene expressions of SOD, GCL, and components of the autophagy signaling pathway. Finally, the performance of r4-Gal4;UAS-AMPKRNAi flies confirmed the regulatory role of the autophagy signaling pathway in mitigating UV irradiation injury.
Preprint
Full-text available
Drosophila melanogaster has a rich repertoire of innate and learned behaviors. Its 100,000-neuron brain is a large but tractable target for comprehensive neural circuit mapping. Only electron microscopy (EM) enables complete, unbiased mapping of synaptic connectivity; however, the fly brain is too large for conventional EM. We developed a custom high-throughput EM platform and imaged the entire brain of an adult female fly. We validated the dataset by tracing brain-spanning circuitry involving the mushroom body (MB), intensively studied for its role in learning. Here we describe the complete set of olfactory inputs to the MB; find a new cell type providing driving input to Kenyon cells (the intrinsic MB neurons); identify neurons postsynaptic to Kenyon cell dendrites; and find that axonal arbors providing input to the MB calyx are more tightly clustered than previously indicated by light-level data. This freely available EM dataset will significantly accelerate Drosophila neuroscience.
Article
Full-text available
Drosophila melanogaster has a rich repertoire of innate and learned behaviors. Its 100,000-neuron brain is a large but tractable target for comprehensive neural circuit mapping. Only electron microscopy (EM) enables complete, unbiased mapping of synaptic connectivity; however, the fly brain is too large for conventional EM. We developed a custom high-throughput EM platform and imaged the entire brain of an adult female fly at synaptic resolution. To validate the dataset, we traced brain-spanning circuitry involving the mushroom body (MB), which has been extensively studied for its role in learning. All inputs to Kenyon cells (KCs), the intrinsic neurons of the MB, were mapped, revealing a previously unknown cell type, postsynaptic partners of KC dendrites, and unexpected clustering of olfactory projection neurons. These reconstructions show that this freely available EM volume supports mapping of brain-spanning circuits, which will significantly accelerate Drosophila neuroscience. Video Abstract
Article
Full-text available
Here, we provide a brief review of the mechanistic connections between immunity and aging—a fundamental biological relationship that remains poorly understood—by considering two intertwined questions: how does aging affect immunity, and how does immunity affect aging? On the one hand, aging contributes to the deterioration of immune function and predisposes the organism to infections (“immuno-senescence”). On the other hand, excessive activation of the immune system can accelerate degenerative processes, cause inflammation and immunopathology, and thus promote aging (“inflammaging”). Interestingly, several recent lines of evidence support the hypothesis that restrained or curbed immune activity at old age (that is, optimized age-dependent immune homeostasis) might actually improve realized immune function and thereby promote longevity. We focus mainly on insights from Drosophila, a powerful genetic model system in which both immunity and aging have been extensively studied, and conclude by outlining several unresolved questions in the field.
Article
Full-text available
Maternal obesity and metabolic diseases are two of the most important potential dangers to offspring, given that impaired offspring may cause deficiencies that impair the adult life and health. This study evaluated the oxidative damage, the enzymatic antioxidant defenses, and the enzymes of fatty acid metabolism, such as Acyl-CoA Synthetase and Acetyl-CoA Synthetase (mRNA expression levels), as well as the modulation of cell stress signaling pathway, as Hsp83, and gene expression and insulin-like peptide DILP6 in Drosophila melanogaster models that received a high fat diet (HFD) (10% and 20% of coconut oil) throughout their development period. After seven days, the progenitor flies were removed and, the remaining eggs were monitored daily, until the eclosion. The descendants were then exposed to a regular diet (RD). The results revealed that the HFD caused a decrease in the proportion of eclosion, lifespan, MTT reduction in mitochondrial enriched fractions, AceCS1 levels, mRNA expression levels (SOD and CAT), and in catalase activity a decrease was only observed in the group that received the highest concentration of coconut oil. In parallel, it was demonstrated an increase in the upregulation of HSP83 mRNA levels, but only when 10% of coconut oil was added, and an increase in glucose and triglyceride levels, as well as in DILP6 mRNA levels in larger concentration of coconut oil tested (20%). In conclusion, flies that have progenitors fed with HFD can develop metabolic dysfunctions, causing oxidative insults, which are involved in the shortening of lifespan. This article is protected by copyright. All rights reserved
Article
Full-text available
The gastrointestinal tract serves as a fast-renewing model for unraveling the multifaceted molecular mechanisms underlying remarkably rapid cell renewal, which is exclusively fueled by a small number of long-lived stem cells and their progeny. Stem cell activity is the best-characterized aspect of mucosal homeostasis in mitotically active tissues, and the dysregulation of regenerative capacity is a hallmark of epithelial immune defects. This dysregulation is frequently associated with pathologies ranging from chronic enteritis to malignancies in humans. Application of the adult Drosophila gastrointestinal tract model in current and future studies to analyze the immuno-physiological aspects of epithelial defense strategies, including stem cell behavior and re-epithelialization, will be necessary to improve our general understanding of stem cell participation in epithelial turnover. In this review, which describes exciting observations obtained from the adult Drosophila gastrointestinal tract, we summarize a remarkable series of recent findings in the literature to decipher the molecular mechanisms through which stem cells respond to nonsterile environments.
Article
Full-text available
Variation in the quality and quantity of nutrition is a major contributor to phenotypic variation in animal populations. Although we know much of how dietary restriction impacts phenotype, and of the molecular-genetic and physiological mechanisms that underlie this response, we know much less of the effects of dietary imbalance. Specifically, although dietary imbalance and restriction both reduce overall body size, it is unclear whether both have the same effect on the size of individual traits. Here, we use the fruit fly Drosophila melanogaster to explore the effect of dietary food versus protein-to-carbohydrate ratio on body proportion and trait size. Our results indicate that body proportion and trait size respond differently to changes in diet quantity (food concentration) versus diet quality (protein-to-carbohydrate ratio), and that these effects are sex specific. While these differences suggest that Drosophila use at least partially distinct developmental mechanisms to respond to diet quality versus quantity, further analysis indicates that the responses can be largely explained by the independent and contrasting effects of protein and carbohydrate concentration on trait size. Our data highlight the importance of considering macronutrient composition when elucidating the effect of nutrition on trait size, at the levels of both morphology and developmental physiology.
Article
Full-text available
Women live on average longer than men but have greater levels of late-life morbidity. We have uncovered a substantial sex difference in the pathology of the aging gut in Drosophila. The intestinal epithelium of the aging female undergoes major deterioration, driven by intestinal stem cell (ISC) division, while lower ISC activity in males associates with delay or absence of pathology, and better barrier function, even at old ages. Males succumb to intestinal challenges to which females are resistant, associated with fewer proliferating ISCs, suggesting a trade-off between highly active repair mechanisms and late-life pathology in females. Dietary restriction reduces gut pathology in aging females, and extends female lifespan more than male. By genetic sex reversal of a specific gut region, we induced female-like aging pathologies in males, associated with decreased lifespan, but also with a greater increase in longevity in response to dietary restriction.
Article
Excess adipose fat accumulation, or obesity, is a growing problem worldwide in terms of both the rate of incidence and the severity of obesity-associated metabolic disease. Adipose tissue evolved in animals as a specialized dynamic lipid storage depot: adipose cells synthesize fat (a process called lipogenesis) when energy is plentiful and mobilize stored fat (a process called lipolysis) when energy is needed. When a disruption of lipid homeostasis favors increased fat synthesis and storage with little turnover owing to genetic predisposition, overnutrition or sedentary living, complications such as diabetes and cardiovascular disease are more likely to arise. The vinegar flyDrosophila melanogaster(Diptera: Drosophilidae) is used as a model to better understand the mechanisms governing fat metabolism and distribution. Flies offer a wealth of paradigms with which to study the regulation and physiological effects of fat accumulation. Obese flies accumulate triacylglycerols in the fat body, an organ similar to mammalian adipose tissue, which specializes in lipid storage and catabolism. Discoveries inDrosophilahave ranged from endocrine hormones that control obesity to subcellular mechanisms that regulate lipogenesis and lipolysis, many of which are evolutionarily conserved. Furthermore, obese flies exhibit pathophysiological complications, including hyperglycemia, reduced longevity and cardiovascular function - similar to those observed in obese humans. Here, we review some of the salient features of the fly that enable researchers to study the contributions of feeding, absorption, distribution and the metabolism of lipids to systemic physiology.
Article
We hypothesized that the administration of explosion-puffed coffee, containing γ-aminobutyric acid (GABA) and 5-hydroxytryptophan (5-HTP), would be associated with a reduction of the caffeine effect on sleep behavior and behavioral patterns, which was investigated in a Drosophila model. The effects of feeding roasted coffee beans (RB), explosion-puffed coffee beans puffed at 0.75. MPa and 0.9. MPa (PB 7.5 and PB 9.0, respectively), or decaffeinated coffee beans (DeRB) on locomotor activity and behavioral patterns of Drosophila was analyzed. In the decreasing order, the total chlorogenic acid (caffeoylquinic acids, CQA) content was PB 7.5. >. PB 9.0. >. RB. PB content analysis showed high levels of GABA and 5-HTP, compared with that of RB, which corresponded with the sleep-wake behavior of Drosophila. The RB and PB (PB 7.5 and PB 9.0) groups were not significantly different with respect to an activity count during the subjective night and day period compared with the normal controls. Sleep bout numbers of the normal, PB, and DeRB groups showed significant differences as compared with the caffeine and RB groups (p <. 0.05). The PB and DePB groups showed a significantly increased transcript levels for the GABA receptors compared to the caffeine group. The caffeine and RB groups displayed better climbing ability than the other groups, covering an average distance 6. cm in the related test; the average distance covered by the normal, PB 7.5, and DeRB groups was <. 4. cm. The normal and DeRB groups showed similar behavior patterns with respect to total distance, velocity, moving, not moving, and meander. However, the PB 7.5 group significantly regulated not moving and meander of flies compared to flies receiving only caffeine and RB. Suppression of the stimulating effect of caffeine by explosion-puffed coffee administration was indicated in the above results, which can be attributed to the increased content of GABA and 5-HTP with explosive puffing process carried out at 0.75. MPa. Results of the underlying mechanism of the behavioral change patterns of explosive puffed with or without caffeine in Drosophila models, transcript level for the Dop1-R1 receptor in caffeine group was significantly higher than normal, PB, and DePB groups. Flies exposed to the caffeine had significantly decreased transcript levels for the GABA receptors. PB 7.5 and DePB showed higher level of GABA content than RB.