Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Nervous systems endow animals with cognition and behavior. To understand how nervous systems control behavior, neural circuits mediating distinct functions need to be identified and characterized. With superior genetic manipulability, Drosophila is a model organism at the leading edge of neural circuit analysis. We briefly introduce the state-of-the-art genetic tools that permit precise labeling of neurons and their interconnectivity and investigating what is happening in the brain of a behaving animal and manipulating neurons to determine how behaviors are affected. Brain-wide wiring diagrams, created by light and electron microscopy, bring neural circuit analysis to a new level and scale. Studies enabled by these tools advances our understanding of the nervous system in relation to cognition and behavior.

Free full text 


Logo of neuroscibullNeuroscience Bulletin
Neurosci Bull. 2019 Dec; 35(6): 1058–1072.
Published online 2019 May 22. https://doi.org/10.1007/s12264-019-00390-9
PMCID: PMC6864024
PMID: 31119647

Recent Advances in the Genetic Dissection of Neural Circuits in Drosophila

Abstract

Nervous systems endow animals with cognition and behavior. To understand how nervous systems control behavior, neural circuits mediating distinct functions need to be identified and characterized. With superior genetic manipulability, Drosophila is a model organism at the leading edge of neural circuit analysis. We briefly introduce the state-of-the-art genetic tools that permit precise labeling of neurons and their interconnectivity and investigating what is happening in the brain of a behaving animal and manipulating neurons to determine how behaviors are affected. Brain-wide wiring diagrams, created by light and electron microscopy, bring neural circuit analysis to a new level and scale. Studies enabled by these tools advances our understanding of the nervous system in relation to cognition and behavior.

Keywords: Drosophila, Neural circuit, Neuroanatomy, Neurogenetics, Systems neuroscience

The nervous system endows an animal with cognition and behavior. To understand how cognitive processes and behavior are coordinated and performed, neural circuits mediating distinct functions need to be identified and characterized. With a nervous system of intermediate complexity, a rich behavioral repertoire, and genetic tractability, Drosophila is a model organism at the leading edge of neural circuit analysis. In this review, we give a brief introduction to the state-of-the-art genetic tools commonly used by Drosophilists to precisely target discrete neuronal populations, to reveal neuronal morphology and connectivity, and to monitor and manipulate neuronal activity. We discuss what brain wide-wiring diagrams created by light and electron microscopy might bring to neural circuit analysis. Modeling and evolutionary comparison of neural circuits could aid the understanding of neural circuits, in addition to the genetic dissection of the molecular and neuronal architectures for cognition and behavior.

Genetic Targeting of Neuron Populations

Genetic dissection of neural circuits in Drosophila involves delineating neuronal subtypes, neuronal interconnectivity and signaling, as well as functions in cognition and behavior. Unlike conventional dissection based solely on anatomical knowledge, genetic dissection relies on genetic tools that allow the expression of effector genes in subsets of neurons, thus affording the targeted labeling and/or manipulation of identified neurons. In flies, these tools are mostly based on binary expression systems. Recently, the binary expression systems have been further improved to allow more precise labeling of neurons, and easier repurposing of drivers.

Genetically Accessing Neurons via Binary Expression Systems

Using the transposable P element to introduce transgenes into Drosophila has revolutionized the genetic study of Drosophila [1]. The P element-mediated transformation with the yeast transcription activator GAL4 revealed that the expression of the foreign gene GAL4 can be controlled by local genomic elements, such as enhancers [2]. Introducing both GAL4 and its binding sequence UAS (Upstream Activating Sequence) into flies led to the invention of the most used binary expression system, GAL4/UAS [3], in which the transgene following UAS was expressed only in the cells expressing GAL4, thus allowing targeted expression of effector genes simply by crossing a GAL4 line with a UAS line (Fig. 1A).

An external file that holds a picture, illustration, etc.
Object name is 12264_2019_390_Fig1_HTML.jpg

Genetic targeting of transgene expression in Drosophila. A GAL4/UAS binary expression system. Expression of the transgene (indicated by dotted region) following the UAS is dictated by GAL4, whose expression is under the control of a regulatory sequence such as a promoter. B GAL4/UAS/GAL80 allow a set-difference expression of transgene. The UAS-transgene is expressed in cells expressing GAL4 controlled by promoter 1 but not GAL80 (which binds to GAL4 and inhibits its action on UAS) controlled by promoter 2. C split-GAL4 allows an intersectional expression of transgene. The transgene following the UAS is expressed in cells expressing both split-GAL4 parts, the GAD and DBD, which are under the control of promoter 1 and promoter 2, respectively. D FLP/FRT adds another layer of control over a binary expression system such as GAL4/UAS. The UAS-transgene is expressed in cells expressing both GAL4 and FLP, which stochastically excise the FRT-stop cassette to allow the effector gene to be expressed.

GAL4 lines were previously constructed with the GAL4 coding sequence following a minimal promotor (enhancer trap) [46] or a defined promoter (enhancer fusion) [3] integrated into the genome with the P element at random loci. Expression of the enhancer trap GAL4 depends on the local cis-regulatory sequences such as enhancers. Recently, the community has been increasingly drawn to newer collections of driver lines such as the Janelia FlyLight [7, 8] and Vienna Tiles (VT) lines [9] that are generated with a modular design: optimized vectors, defined cis-regulatory sequences, and characterized genomic loci for PhiC31-mediated targeted integration. The expression patterns of most FlyLight and VT lines are visualized and searchable, which makes it easier to find useful drivers expressing in brain regions of interest and greatly facilitates the dissection of neural circuits.

Binary expression systems other than GAL4/UAS are also frequently used to allow independent expression of different transgenes in different neuron populations. The LexA/LexAop system introduces both bacterial transcription repressor LexA’s DNA binding domain fusing the viral transcription factor VP16 acidic activation domain (VP16AD) or the GAL4 trans-activating domain (GAD) or human p65 activation domains (p65AD) and LexA binding sequence LexAop into flies [8, 10]. The QF/QUAS system borrows the transcriptional activator QF and its binding sequence QUAS from Neurospora crassa [11, 12].

Spatiotemporally Refining Genetic Targeting

There are about 100, 000 neurons in the fly nervous system. Only in rare cases do drivers label sparse or even a single pair of neurons. Genetic strategies have been developed to narrow down the expression of drivers and make it possible to study brain functions at single-cell resolution.

One strategy takes advantage a repressor of the transcription activator. GAL80 binds to the GAD and inhibits GAL4 action on UAS, thus independently targeting GAL4 and GAL80 allows expression of the UAS-transgene in neurons expressing GAL4 but not GAL80, achieving a narrowed set-difference expression pattern [13] (Fig. 1B). elav-GAL80 [14] represses the expression of GAL4 pan-neuronally and is often used to ascertain that GAL4-driven expression of the effector, in neural tissue indeed, is responsible for the observed effect; tsh-GAL80 [15] can mask the expression of GAL4 in most ventral nerve cord (VNC) neurons, and is useful for differentiating neurons in the brain from those in the VNC in some cases. The temperature-sensitive and ubiquitously expressed GAL80 (tubP-GAL80ts) is used to achieve temporal control over transgene expression by transferring the restrictive 18°C to permissive 30°C to relieve GAL8 repression, e.g., to have transgene expressed only during the adult stage to circumvent developmental effects [3, 16]. In the QF/QS/QUAS tertiary expression system, suppression of QF by its repressor QS can be relieved by feeding quinic acid to flies, allowing spatiotemporal control of the transgene expression [11].

Another strategy takes advantage of reconstitution of a functional transcription factor from its split parts. Split-GAL4 fuses the complementary DNA-binding domain (DBD) and transcription-activation domain GAD to each of a pair of heterodimerizing leucine zippers (Zip and Zip+), to allow them to reconstitute functional GAL4. Transgenes under UAS control are expressed only in cells expressing both DBD and GAD (and thus reconstituted GAL4). Targeting the DBD and GAD independently produces intersectional expression of two drivers [17] (Fig. 1C). The orthogonal modular design of constructs for FlyLight facilitates the generation of split-GAL4 lines with the same regulatory sequences and thus mostly reproducible expression patterns of GAL4 lines. Recently, as more split-GAL4 lines have been characterized [18], systematic dissection of neuronal subtypes in various complex brain areas such as the lamina [19], lobula complex [20], mushroom bodies output neurons [21], and the sexually dimorphic P1 neurons [22] has become feasible. A split-LexA system that independently targets the DNA-binding LexA and trans-activating VP16AD has also been developed to allow intersectional expression [23].

Combining split-GAL4 and GAL80 can further narrow down the expression of transgenes. However, while GAL80 represses GAD-mediated transcription, it does not effectively inhibit split-GAL4 that uses p65AD or VP16AD as transcriptional activators. Killer Zipper (KZip+), a dominant-negative repressor of split-GAL4 that disrupts the formation of the split-GAL4 heterodimer, has been introduced to allow targeted expression in neurons expressing both parts of split-GAL4 but not GAL80 [24].

Stochastic Labeling by Site-Specific Recombination

Recombination-mediated excision of functional cassettes in the transgene adds a separate layer of control over the binary expression systems. The yeast recombinase Flippase (FLP) catalyzes mitotic recombination between the FRT (flippase recognition target) sequences on homologous chromosomal loci. FLP/FRT is introduced into Drosophila [25, 26] for the induction of stochastic labeling by inserting FLP-excisable FRT cassettes into effector constructs such as the UAS-transgene and targeted recombination. Directed flanking FRTs can generate “flip out” excision of the flanked cassette while oppositely-directed flanking FRTs can generate inversion of the flanked cassette. An FRT-flanked cassette with a transcriptional terminator (FRT-stop) inserted after UAS and just before the transgene is often used to “flip in” the transgene in neurons with FLP, which excises the FRT-stop cassette, and narrows down expression of the transgene in neurons intersectionally defined by both independently-targeted GAL4 and FLP [27] (Fig. 1D). For example, FLP under control of the brain-restricted Otd promoter Otd-FLP has been used in combination either with tubP > GAL80 > (> indicates FRT site) or with UAS > stop > genes allows effector transgenes to be expressed in the brain [2830]. Heat-shock-inducible FLP (hs-FLP) allows control over the timing and volume of recombination events which eventually results in varying degrees of labeling. Other site-specific recombination systems have also been introduced into Drosophila, such as Cre/LoxP from bacteriophage [31], KD/KDRT (KD-recombinase Target Recognition) from Kluyveromyces drosophilarum, R/RSRT from Zygosaccharomyces rouxii, and B3/B3RT (B3 recombinase Target Recognition) from Zygosaccharomyces bisporus [32].

MARCM (Mosaic Analysis with a Repressible Cell Marker) uses hs-FLP/FRT and the ubiquitously-expressed GAL80 to induce mosaic cells with transgene expression in targeted tissue without transgene expression [13]. MARCM affords sparse expression of reporters or effectors in a targeted neuron population, and allows study of the morphology and development as well as the behavioral functions of neural circuits at single-cell resolution [33].

FINGR (Flippase-induced Intersectional GAL80/GAL4 Repression) combines the FRT-stop with GAL80 to allow FRT-dependent “flip-in” of GAL80 to repress GAL4 expression with a collection of enhancer-trap Flippases [34] and allows more restricted targeting in neurons with GAL4 but without FLP expression.

Repurposing GAL4 to GAL80, Split-GAL4, and QF2

Tasks such as revealing the relationship between two neuronal populations entail simultaneous targeting of each by independent binary systems. As most available drivers are GAL4 lines, other binary drivers such as LexA and QF that reproduce the GAL4 expression are needed. Therefore, easy methods to repurpose GAL4 lines to other drivers are of great interest to the community. Previous attempts such as InSITE (Integrase Swappable in vivo Targeting Element) build platforms to allow GAL4 be replaced by LexA, QF, or GAL80 by crosses. However, the system can only replace specially-constructed IT-GAL4s but cannot replace existing GAL4s [35]. Recently, the HACK (Homology Assisted CRISPR Knock-in) method was invented to replace GAL4 in the genome with split-GAL4, GAL80 [36], and QF2 [37] by harnessing gene conversion that uses the clustered regularly-interspaced palindromic repeats (CRISPR)/Cas9 system to induce double-strand DNA breaks in the GAL4 sequence and then homology-directed repair to unidirectionally transfer a donor homologous sequence. The FlyLight project allows relatively convenient generation of LexA or QF2 or split-GAL4 lines with largely reproducible expression as GAL4 lines by using the same regulatory sequences [38]. However, the HACK method can convert GAL4 lines only by crosses without plasmid construct and injection, thus is particularly useful to repurpose enhancer trap GAL4 lines whose expression patterns are difficult to reproduce since regulatory sequences are difficult to define.

Revealing Neuronal Morphology by the Expression Pattern of Reporters

Visualizing Neuronal Morphology with Fluorescent Proteins

The expression patterns of drivers can be visualized by light microscopy with the targeted expression of reporters such as fluorescent proteins. Besides the most commonly used green fluorescent protein (GFP) [39], fluorescent proteins with different spectral properties such as red fluorescent proteins (such as tdTomato [40], mCherry [41], and dsRed [42, 43]) and GFP-derived yellow fluorescent proteins (such as Venus[44] and Citrine[45]), and cyan fluorescent proteins (such as Cerulean[46] and mTurquoise[47]) are also frequently used. To reveal different cellular components or compartments, GFP has been engineered by fusing it to various subcellularly localized molecules or by post-translational modification. GFP fused with mCD8 [13] or myristoylated GFP [8] is frequently used to mark cell membranes. GFP tagged with the synaptic vesicle-associated protein Synaptotagmin 1 (Syt::GFP) allow labeling of presynaptic compartments [48]. A hybrid protein of mCherry and the mouse protein Icam5, which is specifically and highly enriched in somatodendritic compartments (DenMark), is used to mark postsynaptic sites [49]. GFP fused with the nuclear localization signal from the large T antigen of Simian virus 40 allows nuclei to be labeled [50] and is useful to discern cell identity and count cell numbers. RedStinger, DsRed(T4) tagged with a nuclear localization signal, is also used to mark nuclei [43] (Table (Table11).

Table 1

Common effectors / methods for labeling, monitoring or manipulating neurons.

Effectors/methodsDescription
Fluorescent reporters
UAS-GFPmCD8::GFP [8] and myr::GFP [39] to mark cell membranes
UAS-RFPtdTomato [40], mCherry [41], dsRed [42, 43]
UAS-YFPVenus [44] and Citrine [45]
UAS-CFPCerulean [46] and mTurquoise [47]
Subcellular marker
UAS-DenMarkMouse Icam5-tagged mCherry to mark the postsynaptic sites [49]
UAS-syt::GFPSyt1-fused GFP to label presynaptic compartments [48]
UAS-GFP::nlsNucleus-localized GFP [50]
UAS-RedStingerNucleus-localized DsRed(T4) [43]
UAS-C3PA-GFPIntracellular diffusion of focused photoactivated PA-GFP to reveal neuronal processes [92, 93]
Genetically-encoded Ca2+indicators
UAS-IVS-GCaMP6*Most-used GFP-based GECIs [56] GCaMP6s, GCaMP6f [58]
UAS-RCaMP*mRuby-based GECIs jRCaMP1a, b [64]
UAS-R-GECO*mApple-based GECIs jRGECO1a, 1b [64]
UAS-twitch-2C*FRET-based ratiometric fluorescent Ca2+ reporter [60, 61]
UAS-CaMPARIRatiometric reporter reflects Ca2+ activity under photoconverting UV [67]
CaLexATranscriptional reporter for sustained Ca2+ activity
TRICTranscriptional reporter for sustained Ca2+ activity
Genetically-encoded voltage indicators
UAS-ArcLightGEVI with improved signal and signal-to-noise ratio [80]
UASp-Asap2f*GEVI with improved sensitivity [81]
Neuronal manipulation
UAS-rprPro-apoptotic molecule Reaper[84] to ablate cells
UAS-hidPro-apoptotic molecule Hid [85] to ablate cells
UAS-TNT*Tetanus toxin light chain to suppress synaptic vesicle fusion and block neurotransmission [87]
UAS-shibirets*Temperature-sensitive dominant-negative of Dynamin to disrupt synaptic vesicle exo/endocytosis and block neurotransmission [88]
UAS-NachBacBacterial Na+ channel to increase neuronal excitability [89]
UAS-Kir2.1*K+ inwardly-rectifying channel to silence neuronal firing [90]
UAS-Ork1.Δ-COpen rectifier K+ channel to silence neuronal firing [91]
UAS-TrpA1Thermosensitive cation channel to activate neurons [94]
UAS-ChR2*Blue light-sensitive Channelrhodopsin to activate neurons [95]
UAS-ReaChR*Red light-sensitive Channelrhodopsin to activate neurons [96]
UAS-CsChrimson*Red light-sensitive Channelrhodopsin to activate neurons [97]
UAS-NpHR.YFPYellow light-sensitive Halorhodopsin to suppress neurons [98100]

UAS-GtACR1

UAS-GtACR2

Cyan (GtACR1) or blue (GtACR2) light-sensitive

Channelrhodopsin to silence neurons [101]

UAS-P2X2.Y*ATP-gated non-specific cation channel to activate neurons [102]

*LexAop lines also available.

Revealing the Composition of Neural Circuits with Stochastic Multi-color Labeling

Plexuses of neurons in neural circuits are often intractable for one fluorescent reporter to discern its neuronal diversity and arrangement. Stochastic multicolor labeling can visualize multiple neurons in distinct colors in the same preparation, and thus allows the tracing of neuronal lineages and projections to reveal how neural circuits are arranged. Brainbow has been developed in the mouse to label neurons with different colors by stochastic expression of one of multiple fluorescent proteins based on Cre/Lox recombination [51]. Brainbow has been introduced into Drosophila using two approaches: Flybow uses a modified inducible FLP/FRT system to drive the inversion and excision of FRT-flanked cassettes of different membrane-tethered fluorescent proteins to allow stochastic expression of one of four epitopes [52] (Fig. 2A); dBrainbow uses Cre and three mutually-exclusive Lox sites to drive the stochastic expression of one of three epitope-tagged proteins, and double copies of dBrainbow can generate six-color labeling [53] (Fig. 2B). MCFO (MultiColor Flip Out), allows multicolor stochastic labeling by controlling the expression of multiple distinct epitope-tagged proteins with recombinase-mediated excision of transcription-terminating cassettes [54] (Fig. 2C). This has been used to reveal the shapes and arrangements of cells in the visual system [20] and central complex [55], among other brain regions.

An external file that holds a picture, illustration, etc.
Object name is 12264_2019_390_Fig2_HTML.jpg

Stochastic multicolor labeling of cells. A Flybow uses hs-FLP (enzyme for FRT), hs-mFlp5 (enzyme for mFRT71) to generate FRT-mediated inversion and excision to allow expression of one of four fluorescent proteins in subsets of targeted neurons. B dBrainbow uses hs-Cre to induce excision of the stop cassettes to allow one of the three fluorescent protein to be stochastically expressed in targeted neurons. Two copies of the constructs generate six-color labeling. C MCFO uses hs-FLP to stochastically excise FRT-flanked interruption cassettes and allow the combinatorial expression of three markers.

Monitoring Neuronal Activity by a Genetically-Encoded Ca2+ Indicator or Voltage Indicator

The functioning of neural circuits is reflected in dynamic and patterned neuronal activity. Ca2+ and voltage dynamics are common indicators of neuronal activity and can be monitored by light microscopy imaging with genetically-encoded Ca2+ indicators (GECIs) and voltage indicators (GEVIs), respectively. Targeted expression of GECIs or GEVIs allows the monitoring of neuronal activity in labeled neurons or neuronal compartments (Table (Table11).

Monitoring Neuronal Activity with Genetically-Encoded Ca2+ Indicators

Intracellular Ca2+ is a widely-used indicator for neuronal activity. Various genetically-encoded indicators have been developed, based on fluorescent proteins and Ca2+-binding proteins such as Calmodulin (CaM), allowing activity to be monitored in a targeted neuronal population.

Monitoring Transient Ca2+ in Neurons

Transient Ca2+ can be monitored using GCaMP, a single molecule GFP-based Ca2+ probe with the CaM and its receptor, the M13 fragment from myosin light chain kinase, attached to the C and N termini of a circularly-permutated EGFP (cpEGFP) [56]. Ca2+-binding to CaM induces a conformational change of CaM-M13 and thus the cpEGFP, resulting in a change in fluorescence intensity. It is widely used to image neuronal responses [57]. The latest version, GCaMP6, can even detect single action potentials in the soma [58]. Subcellularly-targeted GCaMP allows the compartmentalized imaging of Ca2+. Syt::GCaMP, which attaches GCaMP6 to the C terminus of Syt1, then locates to and detects Ca2+ influx specifically in presynaptic terminals [59].

Transient Ca2+ can also be monitored using Twitch-2C, a ratiometric fluorescent Ca2+ reporter that utilizes ‘Twitch’ sensors based on the C-terminal domain of Opsanus troponin C and fluorescence resonance energy transfer (FRET) [60]. A conformational change mediated by Ca2+ binding alters the FRET between two fluorophores (mTurqoise2 and cpCitrine174), resulting in a change in the ratio of cyan to yellow fluorescence, which can be monitored by ratiometric imaging [61].

Red versions of GECIs, such as R-GECO (based on mApple) [62] and RCaMPs (based on mRuby) [63], of which the recent versions jRCaMP1a, b or jRGECO1a [64] have sensitivity comparable to GCaMP6, provide an additional channel of imaging to allow dual-color imaging of separate neuronal populations when used together with GCaMP to simultaneously monitor two populations of neurons [65].

GECIs can be used for neuron imaging in acutely dissected live nervous system or head fixed animal. Flyception allows more temporarily precise real-time imaging in freely-moving flies through an imaging window with the head cuticle removed in a specially-designed arena [66].

Capturing Neuronal Ca2+ Activity in a Light-Defined Time-Window

CaMPARI (calcium-modulated photoactivatable ratiometric integrator) takes advantage of the photoconvertible green fluorescent protein (FP) that irreversibly converts to a red fluorescent species upon illumination with violet light. The construct attaches a circular permutation of a photoconvertible FP to CaM and M13 peptide to allow irreversible conversion from green to red fluorescence with an increase in Ca2+ and the simultaneous presence of UV light (405 nm). It is used to label the increased activity of targeted neurons during behavior by simultaneous UV photoconversion [67]. CaMPARI has recently been used to identify clock neuron classes that respond to changes in temperature [68], and verify the gustatory response of interneurons involved in bitter processing in affixed adult flies [69], and to monitor the response of dopamine neurons to odor [70] and octopaminergic neurons to behavioral interruption training in freely-behaving larvae [71].

Transcriptional Reporter for Ca2+ level

CaLexA (Ca2+-dependent nuclear import of LexA) takes advantage of a Ca2+-responsive transcription factor, nuclear factor of activated T cells, and fuses it to VP16AD and mutant LexA that lacks a nuclear localization signal. Sustained neuronal activation promotes nuclear transport of the LexA chimera and drives transgene expression under the control of LexAop [72].

TRIC (transcriptional reporter of intracellular Ca2+) combines split GAL4 parts with CaM and its targeting peptide CaMKII to reconstitute functional GAL4, contingent upon the Ca2+ level. By fusing GAD to CaM and DBD to CaMKII, Ca2+-binding facilitates the association of CaM with CaMKII and thus the reconstitution and nuclear transport of functional GAL4. Thus, the expression of reporter genes under GAL4/UAS control reflects a sustained neuronal activation [73].

CaLexA and TRIC are both useful for reporting chronic neuronal activation [74, 75]. A real-time CaLexA-LUC assay combining UAS-CaLexA with LexAop-LUC has revealed the activity of circadian neurons in intact freely-moving flies [76]. A similar Tric-LUC has also been developed [77].

Monitoring Neuronal Activity with Genetically-Encoded Voltage Indicators

Electrophysiology is the standard method to monitor the electrical dynamics of a neuron. Whole-cell patch-clamp recording has been increasingly used to monitor central neurons and map neural circuits [78, 79]. The electrophysiological responses of neurons can also be monitored by light microscope imaging with GEVIs, which allow the optical detection of voltage transients at the subcellular level in targeted neurons. ArcLight, combining superecliptic pHlourin and the S1–S4 voltage-sensor domain of the Ciona intestinalis voltage-sensitive phosphatase, can report both subthreshold events and action potentials in intact neurons with an improved signal and signal-to-noise ratio [80]. ASAP is another GEVI with improved sensitivity and allows two-photon imaging of graded potentials in flies [81]. ASAP2f [82], an improvement on ASAP, produces larger fluorescence changes and sustains stable responses during continuous imaging. With superior sensitivity and kinetics, the potential of GEVIs for optical electrophysiology is mostly hindered by that the temporal scanning speed does not match the dynamics of neurons [83].

Neuronal Manipulation

Neurons can be manipulated to interrogate their functions. Ablating neurons genetically by pro-apoptotic molecules such as Reaper [84] or Hid [85], or making neurons dysfunctional by blocking transmission or repressing electrical pulses can reveal the necessity of neurons for behaviors. Activating neurons can uncover behavior-initiating/inducing neurons, like those that control the feeding motor program [86].

Neuronal transmission can be blocked using tetanus toxin light chain (TNT), which suppresses synaptic vesicle fusion and thus neurotransmission [87], or Shibire, a dominant-negative form of Dynamin, an essential component of the synaptic vesicle exo/endocytosis machinery [88]. Neuronal excitability can be promoted by the bacterial depolarization-activated Na+ channel NaChBac [89] or silenced by the inwardly-rectifying K+ channel Kir2.1 [90] or by open rectifier K+ channel dOrk-ΔC [91]. Targeted expression of specific effectors allows neuronal activity to be conditionally modulated by light, temperature, or chemicals (Table 1).

Optogenetics Controls Neuronal Activity with Light

Optogenetics tools use light-sensitive ion channels to achieve manipulation of neuronal activity with high temporal resolution. Neurons can be activated by blue light with Channelrhodopsin-2 (ChR2), a blue light (470 nm)-sensitive nonselective cation channel derived from the green alga Chlamydomonas reinhardtii [95], However, blue light has a visual effect on flies and does not penetrate the cuticle well. The side-effects can be circumvented by using red light. ReaChR (Red-activatable Channelrhodopsin) is derived from ChR1 in Chlamydomonas reinhardtii with a red-shifted spectral sensitivity that allows neuronal activation by red light (590 nm–630 nm) [96]. CsChrimson, derived from red light (590 nm)-responsive ChR1 in the red alga Chlamydomonas noctigama, has been introduced into Drosophila to activate neurons with red light [97].

The Cl channel Halorhodopsin from the archaebacterium Natronomonas pharaonis (NpHR) is able to suppress neuronal firing upon activation by yellow light (570 nm) [98100]. The theta anion Channelrhodopsins from algal Guillardia (GtACRs) are have been introduced into flies for optogenetic silencing of neurons by cyan (GtACR1, at 515 nm) or blue light (GtACR2, at 470 nm) [101].

Thermogenetics Controls Neuronal Activity with Temperature

Neuronal function can be modulated by temperature in the poikilotherm Drosophila with temperature-sensitive molecules. Neuronal transmission can be temporally blocked by temperature-sensitive Shibirets1 at temperatures above 29°C [88]. Neurons can be activated by the thermosensitive cation channel TrpA1 if shifted from the restrictive 22°C or below to above 25°C [94].

Chemogenetics Controls Neuronal Activity with Chemicals

The ATP (adenosine 5’-triphosphate)-gated non-specific cation channel receptor P2X2 was used in flies as the first demonstration of controlling neuronal activity by light [102]. It was initially used in flies for light activation of neurons by injecting the flies with encased ATP which can be uncaged by light. Drosophila lacks endogenous ATP-gated channels and thus the released ATP can specifically activate neurons expressing P2X2.

Revealing Trans-synaptic Neuronal Connections

Neurons function in circuits. Understanding neuronal function entails identifying the neuronal projections, connecting neurons, and synaptic contacts.

Tracing Neuronal Processes

Neuronal projection of sparsely labeled neurons can be easily visualized by fluorescent reporter. However, neuronal projection within a neural plexus is difficult to discern, which can be relieved by the PA-GFP (photoactivatable GFP), a GFP variant that greatly increases its fluorescence with 488 nm excitation light after intense 413 nm radiation. With targeted expression of PA-GFP in neurons [103], light treatment at specific loci harboring part of the neuronal processes can convert the PA-GFP to its high-fluorescence form and diffusion of the converted GFP allows the projection pattern of a subset of targeted neurons to be highlighted [92, 93]. Although it is invasive to achieve precise light treatment, PA-GFP allows higher spatial specificity than that permitted by genetic targeting, and can reveal connections to discrete brain regions using a broadly-expressed PA-GFP [104].

Trans-neuronal Tracing

To reveal the circuitry in which neurons perform functions, it is necessary to map the downstream and upstream neurons. The relationship between two populations of neurons can be grossly addressed by simultaneously expressing different markers in each to see if overlap or contiguity exists in their expression patterns. To systematically map trans-synaptic connected neurons, new strategies have been developed to label the downstream neurons of target neurons.

trans-Tango allows anterograde trans-synaptic tracing, based on the Tango assay in which a synthetic signaling pathway consisting of two fusion proteins converts the activation of a cell-surface receptor such as G protein-coupled receptor to reporter expression via site-specific proteolysis. With trans-Tango, ligands presented at presynaptic sites in targeted neurons (labeled by UAS-myrGFP) trans-synaptically activate receptors and release membrane-tethered QF to drive expression of the QUAS-controlled reporter in postsynaptic neurons via ubiquitously-expressed Tango that uses the QF/QUAS to drive a marker (QUAS-mtdTomato) in the receptor cells [105] (Fig. 3a). This has been used to trace the APDN-TuBusup-EB circuit that regulates sleep-wake arousal [65, 106], and the connections revealed by trans-Tango can be functionally verified with UAS-driven P2X2 and QUAS-driven GCaMP [65].

An external file that holds a picture, illustration, etc.
Object name is 12264_2019_390_Fig3_HTML.jpg

Methods for anterograde trans-synaptic tracing. A Trans-Tango allows anterograde tracing based on pan-neuronally-expressed Tango of human glucagon (hGCG) and presynaptically targeted expression of the glucagon. The trans-Tango is composed of two chimera proteins: pan-neuronally expressed hGCGR::TEVcs::QF tethers the transcription activator QF to the membrane via linking it to the human glucagon receptor (hGCGR) with a linker containing the cleaving site for N1a protease from the tobacco etch virus (TEVcs); hArr::TEV allows recruitment of TEV to activated hGCGR via human β-arrestin2, and one QUAS reporter. Overall, GAL4/UAS targeted expression of hGCG::hICAM1::dNRXN1 presents hGCG at presynaptic sites via the synaptic protein Neurexin1 (dNRXN1) and extracellular domain of human cell adhesion molecule (ICAM1), and activates hGCGR in postsynaptic neurons and recruits hArr::TEV to cleave QF off the membrane to translocate into the nucleus and initiates reporter expression under QUAS control. B TRACT allows anterograde tracing of neuronal circuitry by ligand-induced intramembrane proteolysis. TRACT is composed of CD19 presynaptically localized with nSyb in neurons targeted by LexA/LexAop, and the pan-neuronally-expressed fusion protein nlgSNTG4 that tethers transcription factor GAL4 and CD19 antibody ID3 to the synaptic sites of cell membranes via the intracellular domain of neuroligin (NLGN), the Notch regulatory region (NRR) and transmembrane domain of Notch. The presynaptic ligand CD19 binding to postsynaptic receptors ID3 partially unfolds the NRR to allow for cleavage of the S2 site by endogenous metalloproteases, and further allows cleavage (S3) by γ-secretase-mediated intramembrane proteolysis of the intracellular domain containing GAL4 which then translocates to the nucleus to activate the transcription of reporter genes such as GFP under UAS control.

TRACT (TRAnsneuronal Control of Transcription) can also label anterograde trans-synaptic neurons based on ligand-induced intramembrane proteolysis. Drosophila n-Synaptobrevin (n-Syb)-fused artificial ligands in LexA-labeled neurons are presented on the presynaptic membrane, and the trans-synaptic interaction with an artificial receptor induces intramembrane proteolysis in post-synaptic neurons to liberate a membrane-bound GAL4 variant, and ultimately drives the expression of a UAS-controlled reporter [107] (Fig. 3B). In principle, combined use of TRACT and trans-Tango allows simultaneous labeling of the primary (UAS-driven reporter) and secondary (QUAS-riven reporter) postsynaptic neurons of a LexA driver.

Revealing Potential Synaptic Contacts Between Neuronal Populations

As synapses are beyond the resolution of the ordinary light microscope, apparent contacts of neurons are not necessarily synaptic connections, so genetic methods have been developed to aid in marking putative synaptic connections.

GRASP (GFP Reconstitution Across Synaptic Partners) makes use of the trans-cellular complementation of split GFP parts (spGFP1-10 and spGFP11) to detect the membrane proximity of two cells. The spGFPs, neither of which is fluorescent alone, are separately expressed in the outer membranes of two populations of cells. If two adjacent cells are in enough proximity, the spGFP could reconstitute into a functional fluorescent GFP, revealing potential synaptic contacts (Fig. 4A). The initial version of GRASP used CD4 to present spGFPs to all the outer membranes of cells [108]. To improve the specificity of synaptic labeling, one of the two spGFPs was anchored to synaptic molecules to target synaptic sites. spGFP1-10 was fused to the transmembrane synapse protein Neurexin for better synaptic targeting [109]. Fusing one of the spGFPs to the C terminus of n-Syb (syb:GRASP) allows directional and activity-dependent GRASP labeling, since the presynaptic vesicle membrane n-Syb can only present the spGFP to the synaptic cleft after vesicle fusion to the cell membrane [110]. Recently, t-GRASP (targeted GFP Reconstitution Across Synaptic Partners) tested different pre-synaptic and post-synaptic/dendritic proteins for targeting both spGFPs, and further improved the synaptic specificity [111].

An external file that holds a picture, illustration, etc.
Object name is 12264_2019_390_Fig4_HTML.jpg

Synaptic labeling. A GRASP detects membrane proximity permitting synaptic contacts between two neurons by the fluorescence of a reconstituted functional GFP from complementary spGFPs expressed in each of the neurons. B STaR allows Brp-fused tags with V5 expressed in FLP-mediated stochastically-labeled neurons (with myr-tdTomato) to mark the active zone, mimicking the endogenous pattern of Brp in a cell-type-specific fashion.

STaR (Synaptic Tagging with Recombination) allows inducible tagging of synaptic proteins to visualize synapses in the processes of sparsely-identified neurons, to aid in identifying synapses for specific neurons using ordinary fluorescent light microscope. It is hardly possible to discern dense synaptic puncta in neuronal processes of small size using antibodies against synaptic proteins, which are usually broadly expressed. STaR uses bacterial artificial chromosomes to harbor an engineered genomic region of synaptic protein with a tag sequence separated by a recombination recognition sequence flanked by a stop sequence that can be excised by recombinase, so the induced expression of tagged synaptic proteins display a level and spatiotemporal pattern similar to that of the endogenous synaptic protein. STaR of Brp, a component protein of presynaptic active zones, with a V5 tag, allows FLP-mediated stochastic expression of the V5-tagged Brp to label active zones [112] in defined neurons that are labeled with the LexAop-driven transgene with a co-translating 2A-LexA cassette [113] downstream of the V5 tag (Fig. 4B). Using GFP instead of V5 also allows labeling of active zones of specific neurons in live animals, and can be used to monitor structural synaptic plasticity [114116]. STaR also allows co-labeling of presynaptic protein and postsynaptic receptors, thus revealing the contact sites of synaptic partners [112].

Functional Neuronal Interconnectivity

Synaptic connections between neurons, as indicated by membrane proximity via GRASP, or reconstructed from electron microscopic volumes are not necessarily functional. Functional connections between neurons are assayed by manipulating one neuron and simultaneously monitoring the other, and it is expected that proper activation of the upstream neurons would induce a delayed but timed response in downstream neurons. The most precise method to date uses paired electrophysiological recording electrodes from two neurons to record one neuron’s response to stimulation of the other. Other methods can also do the trick, e.g., one type of neuron can be activated optogenetically by light with CsChrimson [102, 117, 118] or ReaChR [119], or chemogenetically by applying ATP to the acutely dissected brain with P2X2 [76, 115, 120]; meanwhile, activity of the other type can be monitored by electrophysiological recording [121] or imaging with GCaMP. Such assays could also reveal indirect functional connections or population responses.

The Connectome Approach to Neural Circuits

Brain-Wide Circuit Reconstructed by Light Microscopy

The whole-brain wiring diagram, reconstructed by aligning tens of thousands of stochastically-labeled single neurons into a standard brain, provides a blueprint to evaluate the organization of the brain and to infer how different brain regions are interconnected [122]. This single-cell resolution brain reconstruction also aids in interrogating neuron interconnectivity and producing hypotheses of information flow [123, 124]. NBLAST allows identified neurons be fitted to a standard brain and gives hints about their identity and connections [125] as well as similarity [126] based on morphology.

Harnessing machine learning for behavioral classification and big data computing, brain/behavior correlation has been mapped by large-scale analysis of the expression patterns of neurons and the corresponding behaviors induced by their activation in Drosophila larvae [127] and adults [128]. Such approaches produce a behavior reference atlas for neuronal subsets, providing hints on brain regions that might be implicated specific behaviors.

Brain-Wide Circuit Reconstructed by Electron Microscopy

Electron microscopy (EM) affords synaptic resolution circuit reconstruction [129], and is the gold standard for circuit mapping. With rapid improvements in serial-section EM and data acquisition/processing pipeline, EM volumes and reconstructions of more and more parts of the brain have been obtained. The connectome of the fly visual circuitry [130, 131], the EM volume of the entire nervous system of the Drosophila larva [132], the complete wiring diagram of the Drosophila larval antennal lobe [133], mushroom bodies [134], adult olfactory receptor neurons, and projection neurons [135] have been completed. Recently, an EM volume of the whole adult fly brain was obtained, and reconstruction of the mushroom bodies based on the dataset revealed new synaptic partners [136]. More neural circuits and interconnectivity are expected to be reconstructed from this dataset, culminating with a brain-wide connectome, all neurons and their synaptic connections in Drosophila brain. Despite the challenges ahead in processing and annotating the daunting volume of EM data, a post-connectome era is heralded for neural circuit studies in Drosophila.

EM reconstruction can comprehensively map neuronal types and connectivity, reveal interesting but previously unidentified neurons and connections, shedding light on novel circuit motifs, produce hypotheses at the circuit level that can be cross-validated with results obtained by light microscopy, and provide a basic reference for Drosophila neural circuit studies. Considering that the operation of neural circuits can be dynamically reconfigured by neuromodulators [137], combining EM results with the precise genetic manipulation of neural circuits and revealing behavioral assays could help elucidate the operations and functions of neural circuits.

Concluding Remarks and Perspectives

Even as humble as a fruit fly, it must respond to external visual, chemical, mechanosensory stimuli, find food and mates, choose sites for oviposition, modify its behavior with past experiences. How does the tiny fly achieve the feats with circa 1000, 1000 neurons? Here we have briefly reviewed the common genetic tools currently used in Drosophila to solve such mysteries. Brain functions are executed in neural circuits. The sophisticated array of genetic tools available makes flies one of the best organisms for neural circuit cracking. With precise monitoring and manipulation of the nervous system possible at single-cell resolution, it’s possible to establish causal relationship between neurons and behavior and gain mechanistic understanding of the brain in relation to cognition and behaviors.

Furthermore, it’s instructive to make sense of neural circuits and behavior from modeling and evolutionary perspectives.

Theoretical modeling provides explainable systems with functionally-defined components. The predictive power of modeling guides functional studies of circuits and behavior. The Hassenstein–Reichardt elementary motion detector (EMD) has been proposed as a model for motion vision [138], and it has guided efforts to define the visual circuit implementing the EMD [130, 139]. Recent studies revealed a neural circuit for heading orientation, elegantly fitted by the ring attractor model [120, 140].

Insect brains share a common layout, but also differ [141]. How the sameness and differences in brain and behavior arise in evolutionarily close species provides a window for studying the conservation and variation of neural circuits. With the annotated genomes of non-model organisms and CRISPR/Cas9 for convenient genome editing, neuroethologists can draw more on neurogenetic tools. Transferring Drosophila circuit study techniques to non-model organism [142] would reveal the evolution of neural circuits [143], in adaption to particular needs in related species [144].

Acknowledgements

This review was supported by the National Natural Science Foundation of China (6531000063, 31571093, 31622028, 31471063, and 31671074), the Science Foundation of Jiangsu Province of China (BK20160025), and Fundamental Research Funds for the Central Universities, China (2242016R20028 and 2017FZA7003).

Conflict of interest

All authors claim that there are no conflicts of interest.

References

1. Rubin GM, Spradling AC. Genetic transformation of Drosophila with transposable element vectors. Science. 1982;218:348–353. [Abstract] [Google Scholar]
2. O’Kane CJ, Gehring WJ. Detection in situ of genomic regulatory elements in Drosophila. Proc Natl Acad Sci U S A. 1987;84:9123–9127. [Europe PMC free article] [Abstract] [Google Scholar]
3. Brand AH, Perrimon N. Targeted gene-expression as a means of altering cell fates and generating dominant phenotypes. Development. 1993;118:401–415. [Abstract] [Google Scholar]
4. Manseau L, Baradaran A, Brower D, Budhu A, Elefant F, Phan H, et al. GAL4 enhancer traps expressed in the embryo, larval brain, imaginal discs, and ovary of Drosophila. Dev Dyn. 1997;209:310–322. [Abstract] [Google Scholar]
5. Spradling AC, Stern D, Beaton A, Rhem EJ, Laverty T, Mozden N, et al. The Berkeley Drosophila Genome Project gene disruption project: single P-element insertions mutating 25% of vital Drosophila genes. Genetics. 1999;153:135–177. [Europe PMC free article] [Abstract] [Google Scholar]
6. Hayashi S, Ito K, Sado Y, Taniguchi M, Akimoto A, Takeuchi H, et al. GETDB, a database compiling expression patterns and molecular locations of a collection of GAL4 enhancer traps. Genesis. 2002;34:58–61. [Abstract] [Google Scholar]
7. Pfeiffer BD, Jenett A, Hammonds AS, Ngo TT, Misra S, Murphy C, et al. Tools for neuroanatomy and neurogenetics in Drosophila. Proc Natl Acad Sci U S A. 2008;105:9715–9720. [Europe PMC free article] [Abstract] [Google Scholar]
8. Pfeiffer BD, Ngo TT, Hibbard KL, Murphy C, Jenett A, Truman JW, et al. Refinement of tools for targeted gene expression in Drosophila. Genetics. 2010;186:735–755. [Europe PMC free article] [Abstract] [Google Scholar]
9. Tirian L, Dickson B. The VT GAL4, LexA, and split-GAL4 driver line collections for targeted expression in the Drosophila nervous system. bioRxiv. 2017 10.1101/198648. [CrossRef] [Google Scholar]
10. Lai SL, Lee T. Genetic mosaic with dual binary transcriptional systems in Drosophila. Nat Neurosci. 2006;9:703–709. [Abstract] [Google Scholar]
11. Potter CJ, Tasic B, Russler EV, Liang L, Luo L. The Q system: a repressible binary system for transgene expression, lineage tracing, and mosaic analysis. Cell. 2010;141:536–548. [Europe PMC free article] [Abstract] [Google Scholar]
12. Riabinina O, Luginbuhl D, Marr E, Liu S, Wu MN, Luo L, et al. Improved and expanded Q-system reagents for genetic manipulations. Nat Methods. 2015;12:219–222. [Europe PMC free article] [Abstract] [Google Scholar]
13. Lee T, Luo L. Mosaic analysis with a repressible cell marker for studies of gene function in neuronal morphogenesis. Neuron. 1999;22:451–461. [Abstract] [Google Scholar]
14. Yang CH, Rumpf S, Xiang Y, Gordon MD, Song W, Jan LY, et al. Control of the postmating behavioral switch in Drosophila females by internal sensory neurons. Neuron. 2009;61:519–526. [Europe PMC free article] [Abstract] [Google Scholar]
15. Clyne JD, Miesenbock G. Sex-specific control and tuning of the pattern generator for courtship song in Drosophila. Cell. 2008;133:354–363. [Abstract] [Google Scholar]
16. McGuire SE, Le PT, Osborn AJ, Matsumoto K, Davis RL. Spatiotemporal rescue of memory dysfunction in Drosophila. Science. 2003;302:1765–1768. [Abstract] [Google Scholar]
17. Luan H, Peabody NC, Vinson CR, White BH. Refined spatial manipulation of neuronal function by combinatorial restriction of transgene expression. Neuron. 2006;52:425–436. [Europe PMC free article] [Abstract] [Google Scholar]
18. Dionne H, Hibbard KL, Cavallaro A, Kao JC, Rubin GM. Genetic reagents for making split-GAL4 lines in Drosophila. Genetics. 2018;209:31–35. [Europe PMC free article] [Abstract] [Google Scholar]
19. Tuthill JC, Nern A, Holtz SL, Rubin GM, Reiser MB. Contributions of the 12 neuron classes in the fly lamina to motion vision. Neuron. 2013;79:128–140. [Europe PMC free article] [Abstract] [Google Scholar]
20. Wu M, Nern A, Williamson WR, Morimoto MM, Reiser MB, Card GM, et al. Visual projection neurons in the Drosophila lobula link feature detection to distinct behavioral programs. Elife. 2016;5:e21022. [Europe PMC free article] [Abstract] [Google Scholar]
21. Aso Y, Sitaraman D, Ichinose T, Kaun KR, Vogt K, Belliart-Guerin G, et al. Mushroom body output neurons encode valence and guide memory-based action selection in Drosophila. Elife. 2014;3:e04580. [Europe PMC free article] [Abstract] [Google Scholar]
22. Zhang W, Guo C, Chen D, Peng Q, Pan Y. Hierarchical control of Drosophila sleep, courtship, and feeding behaviors by male-specific P1 Neurons. Neurosci Bull. 2018;34:1105–1110. [Europe PMC free article] [Abstract] [Google Scholar]
23. Ting CY, Gu S, Guttikonda S, Lin TY, White BH, Lee CH. Focusing transgene expression in Drosophila by coupling GAL4 with a novel split-LexA expression system. Genetics. 2011;188:229–233. [Europe PMC free article] [Abstract] [Google Scholar]
24. Dolan MJ, Luan H, Shropshire WC, Sutcliffe B, Cocanougher B, Scott RL, et al. Facilitating neuron-specific genetic manipulations in Drosophila melanogaster using a split GAL4 repressor. Genetics. 2017;206:775–784. [Europe PMC free article] [Abstract] [Google Scholar]
25. Golic KG, Lindquist S. The FLP recombinase of yeast catalyzes site-specific recombination in the Drosophila genome. Cell. 1989;59:499–509. [Abstract] [Google Scholar]
26. Xu T, Rubin GM. Analysis of genetic mosaics in developing and adult Drosophila tissues. Development. 1993;117:1223–1237. [Abstract] [Google Scholar]
27. Smith HK, Roberts IJ, Allen MJ, Connolly JB, Moffat KG, O’Kane CJ. Inducible ternary control of transgene expression and cell ablation in Drosophila. Dev Genes Evol. 1996;206:14–24. [Abstract] [Google Scholar]
28. Asahina K, Watanabe K, Duistermars BJ, Hoopfer E, Gonzalez CR, Eyjolfsdottir EA, et al. Tachykinin-expressing neurons control male-specific aggressive arousal in Drosophila. Cell. 2014;156:221–235. [Europe PMC free article] [Abstract] [Google Scholar]
29. Rezaval C, Pattnaik S, Pavlou HJ, Nojima T, Bruggemeier B, D’Souza LAD, et al. Activation of latent courtship circuitry in the brain of Drosophila females induces male-like behaviors. Curr Biol. 2016;26:2508–2515. [Europe PMC free article] [Abstract] [Google Scholar]
30. Koganezawa M, Kimura K, Yamamoto D. The neural circuitry that functions as a switch for courtship versus aggression in Drosophila males. Curr Biol. 2016;26:1395–1403. [Abstract] [Google Scholar]
31. Siegal ML, Hartl DL. Transgene coplacement and high efficiency site-specific recombination with the Cre/loxP system in Drosophila. Genetics. 1996;144:715–726. [Europe PMC free article] [Abstract] [Google Scholar]
32. Nern A, Pfeiffer BD, Svoboda K, Rubin GM. Multiple new site-specific recombinases for use in manipulating animal genomes. Proc Natl Acad Sci U S A. 2011;108:14198–14203. [Europe PMC free article] [Abstract] [Google Scholar]
33. Kohatsu S, Koganezawa M, Yamamoto D. Female contact activates male-specific interneurons that trigger stereotypic courtship behavior in Drosophila. Neuron. 2011;69:498–508. [Abstract] [Google Scholar]
34. Bohm RA, Welch WP, Goodnight LK, Cox LW, Henry LG, Gunter TC, et al. A genetic mosaic approach for neural circuit mapping in Drosophila. Proc Natl Acad Sci U S A. 2010;107:16378–16383. [Europe PMC free article] [Abstract] [Google Scholar]
35. Gohl DM, Silies MA, Gao XJ, Bhalerao S, Luongo FJ, Lin CC, et al. A versatile in vivo system for directed dissection of gene expression patterns. Nat Methods. 2011;8:231–237. [Europe PMC free article] [Abstract] [Google Scholar]
36. Xie T, Ho MCW, Liu Q, Horiuchi W, Lin CC, Task D, et al. A genetic toolkit for dissecting dopamine circuit function in Drosophila. Cell Rep. 2018;23:652–665. [Europe PMC free article] [Abstract] [Google Scholar]
37. Lin CC, Potter CJ. Editing transgenic DNA components by inducible gene replacement in Drosophila melanogaster. Genetics. 2016;203:1613–1628. [Europe PMC free article] [Abstract] [Google Scholar]
38. Jenett A, Rubin GM, Ngo TT, Shepherd D, Murphy C, Dionne H, et al. A GAL4-driver line resource for Drosophila neurobiology. Cell Rep. 2012;2:991–1001. [Europe PMC free article] [Abstract] [Google Scholar]
39. Yeh E, Gustafson K, Boulianne GL. Green fluorescent protein as a vital marker and reporter of gene-expression in Drosophila. Proc Natl Acad Sci U S A. 1995;92:7036–7040. [Europe PMC free article] [Abstract] [Google Scholar]
40. Shaner Nathan C, Campbell Robert E, Steinbach Paul A, Giepmans Ben N G, Palmer Amy E, Tsien Roger Y. Improved monomeric red, orange and yellow fluorescent proteins derived from Discosoma sp. red fluorescent protein. Nature Biotechnology. 2004;22(12):1567–1572. [Abstract] [Google Scholar]
41. Shearin HK, Macdonald IS, Spector LP, Stowers RS. Hexameric GFP and mCherry reporters for the Drosophila GAL4, Q, and LexA transcription systems. Genetics. 2014;196:951–960. [Europe PMC free article] [Abstract] [Google Scholar]
42. Verkhusha VV, Otsuna H, Awasaki T, Oda H, Tsukita S, Ito K. An enhanced mutant of red fluorescent protein DsRed for double labeling and developmental timer of neural fiber bundle formation. J Biol Chem. 2001;276:29621–29624. [Abstract] [Google Scholar]
43. Barolo S, Castro B, Posakony JW. New Drosophila transgenic reporters: insulated P-element vectors expressing fast-maturing RFP. BioTechniques. 2004;36:436–442. [Abstract] [Google Scholar]
44. Nagai T, Ibata K, Park ES, Kubota M, Mikoshiba K, Miyawaki A. A variant of yellow fluorescent protein with fast and efficient maturation for cell-biological applications. Nat Biotechnol. 2002;20:87–90. [Abstract] [Google Scholar]
45. Griesbeck Oliver, Baird Geoffrey S., Campbell Robert E., Zacharias David A., Tsien Roger Y. Reducing the Environmental Sensitivity of Yellow Fluorescent Protein. Journal of Biological Chemistry. 2001;276(31):29188–29194. [Abstract] [Google Scholar]
46. Rizzo MA, Springer GH, Granada B, Piston DW. An improved cyan fluorescent protein variant useful for FRET. Nat Biotechnol. 2004;22:445–449. [Abstract] [Google Scholar]
47. Goedhart J, van Weeren L, Hink MA, Vischer NO, Jalink K, Gadella TW., Jr Bright cyan fluorescent protein variants identified by fluorescence lifetime screening. Nat Methods. 2010;7:137–139. [Abstract] [Google Scholar]
48. Zhang YQ, Rodesch CK, Broadie K. Living synaptic vesicle marker: Synaptotagmin-GFP. Genesis. 2002;34:142–145. [Abstract] [Google Scholar]
49. Nicolai LJ, Ramaekers A, Raemaekers T, Drozdzecki A, Mauss AS, Yan J, et al. Genetically encoded dendritic marker sheds light on neuronal connectivity in Drosophila. Proc Natl Acad Sci U S A. 2010;107:20553–20558. [Europe PMC free article] [Abstract] [Google Scholar]
50. Shiga Y, Tanaka-Matakatsu M, Hayashi S. A nuclear GFP beta-galactosidase fusion protein as a marker for morphogenesis in living Drosophila. Dev Growth Differ. 1996;38:99–106. [Google Scholar]
51. Livet J, Weissman TA, Kang H, Draft RW, Lu J, Bennis RA, et al. Transgenic strategies for combinatorial expression of fluorescent proteins in the nervous system. Nature. 2007;450:56–62. [Abstract] [Google Scholar]
52. Hadjieconomou D, Rotkopf S, Alexandre C, Bell DM, Dickson BJ, Salecker I. Flybow: genetic multicolor cell labeling for neural circuit analysis in Drosophila melanogaster. Nat Methods. 2011;8:260–266. [Abstract] [Google Scholar]
53. Hampel S, Chung P, McKellar CE, Hall D, Looger LL, Simpson JH. Drosophila Brainbow: a recombinase-based fluorescence labeling technique to subdivide neural expression patterns. Nat Methods. 2011;8:253–259. [Europe PMC free article] [Abstract] [Google Scholar]
54. Nern A, Pfeiffer BD, Rubin GM. Optimized tools for multicolor stochastic labeling reveal diverse stereotyped cell arrangements in the fly visual system. Proc Natl Acad Sci U S A. 2015;112:E2967–E2976. [Europe PMC free article] [Abstract] [Google Scholar]
55. Wolff T, Rubin GM. Neuroarchitecture of the Drosophila central complex: A catalog of nodulus and asymmetrical body neurons and a revision of the protocerebral bridge catalog. J Comp Neurol. 2018;526:2585–2611. [Europe PMC free article] [Abstract] [Google Scholar]
56. Nakai J, Ohkura M, Imoto K. A high signal-to-noise Ca2+ probe composed of a single green fluorescent protein. Nat Biotechnol. 2001;19:137–141. [Abstract] [Google Scholar]
57. Wang JW, Wong AM, Flores J, Vosshall LB, Axel R. Two-photon calcium imaging reveals an odor-evoked map of activity in the fly brain. Cell. 2003;112:271–282. [Abstract] [Google Scholar]
58. Chen TW, Wardill TJ, Sun Y, Pulver SR, Renninger SL, Baohan A, et al. Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature. 2013;499:295–300. [Europe PMC free article] [Abstract] [Google Scholar]
59. Cohn R, Morantte I, Ruta V. Coordinated and compartmentalized neuromodulation shapes sensory processing in Drosophila. Cell. 2015;163:1742–1755. [Europe PMC free article] [Abstract] [Google Scholar]
60. Thestrup T, Litzlbauer J, Bartholomaus I, Mues M, Russo L, Dana H, et al. Optimized ratiometric calcium sensors for functional in vivo imaging of neurons and T lymphocytes. Nat Methods. 2014;11:175–182. [Abstract] [Google Scholar]
61. Schnaitmann C, Haikala V, Abraham E, Oberhauser V, Thestrup T, Griesbeck O, et al. Color processing in the early visual system of Drosophila. Cell. 2018;172:318–330. [Abstract] [Google Scholar]
62. Zhao Y, Araki S, Wu J, Teramoto T, Chang YF, Nakano M, et al. An expanded palette of genetically encoded Ca2+ indicators. Science. 2011;333:1888–1891. [Europe PMC free article] [Abstract] [Google Scholar]
63. Akerboom J, Carreras Calderon N, Tian L, Wabnig S, Prigge M, Tolo J, et al. Genetically encoded calcium indicators for multi-color neural activity imaging and combination with optogenetics. Front Mol Neurosci. 2013;6:2. [Europe PMC free article] [Abstract] [Google Scholar]
64. Dana H, Mohar B, Sun Y, Narayan S, Gordus A, Hasseman JP, et al. Sensitive red protein calcium indicators for imaging neural activity. Elife. 2016;5:e12727. [Europe PMC free article] [Abstract] [Google Scholar]
65. Guo F, Holla M, Diaz MM, Rosbash M. A circadian output circuit controls sleep-wake arousal in Drosophila. Neuron. 2018;100:624–635. [Abstract] [Google Scholar]
66. Grover D, Katsuki T, Greenspan RJ. Flyception: imaging brain activity in freely walking fruit flies. Nat Methods. 2016;13:569–572. [Abstract] [Google Scholar]
67. Fosque BF, Sun Y, Dana H, Yang CT, Ohyama T, Tadross MR, et al. Labeling of active neural circuits in vivo with designed calcium integrators. Science. 2015;347:755–760. [Abstract] [Google Scholar]
68. Yadlapalli S, Jiang C, Bahle A, Reddy P, Meyhofer E, Shafer OT. Circadian clock neurons constantly monitor environmental temperature to set sleep timing. Nature. 2018;555:98–102. [Abstract] [Google Scholar]
69. Bohra AA, Kallman BR, Reichert H, VijayRaghavan K. Identification of a single pair of interneurons for bitter taste processing in the Drosophila brain. Curr Biol. 2018;28:847–858. [Abstract] [Google Scholar]
70. Pu Y, Palombo MMM, Shen P. Contribution of DA signaling to appetitive odor perception in a Drosophila model. Sci Rep. 2018;8:5978. [Europe PMC free article] [Abstract] [Google Scholar]
71. Fei Y, Zhu D, Sun Y, Gong C, Huang S, Gong Z. Repeated failure in reward pursuit alters innate Drosophila larval behaviors. Neurosci Bull. 2018;34:901–911. [Europe PMC free article] [Abstract] [Google Scholar]
72. Masuyama K, Zhang Y, Rao Y, Wang JW. Mapping neural circuits with activity-dependent nuclear import of a transcription factor. J Neurogenet. 2012;26:89–102. [Europe PMC free article] [Abstract] [Google Scholar]
73. Gao XJ, Riabinina O, Li J, Potter CJ, Clandinin TR, Luo L. A transcriptional reporter of intracellular Ca2+ in Drosophila. Nat Neurosci. 2015;18:917–925. [Europe PMC free article] [Abstract] [Google Scholar]
74. Li Q, Gong Z. Cold-sensing regulates Drosophila growth through insulin-producing cells. Nat Commun. 2015;6:10083. [Europe PMC free article] [Abstract] [Google Scholar]
75. Jang YH, Chae HS, Kim YJ. Female-specific myoinhibitory peptide neurons regulate mating receptivity in Drosophila melanogaster. Nat Commun. 2017;8:1630. [Europe PMC free article] [Abstract] [Google Scholar]
76. Guo F, Yu J, Jung HJ, Abruzzi KC, Luo W, Griffith LC, et al. Circadian neuron feedback controls the Drosophila sleep-activity profile. Nature. 2016;536:292–297. [Europe PMC free article] [Abstract] [Google Scholar]
77. Guo F, Chen X, Rosbash M. Temporal calcium profiling of specific circadian neurons in freely moving flies. Proc Natl Acad Sci U S A. 2017;114:E8780–E8787. [Europe PMC free article] [Abstract] [Google Scholar]
78. Wilson RI, Turner GC, Laurent G. Transformation of olfactory representations in the Drosophila antennal lobe. Science. 2004;303:366–370. [Abstract] [Google Scholar]
79. Maimon G, Straw AD, Dickinson MH. Active flight increases the gain of visual motion processing in Drosophila. Nat Neurosci. 2010;13:393–399. [Abstract] [Google Scholar]
80. Cao G, Platisa J, Pieribone VA, Raccuglia D, Kunst M, Nitabach MN. Genetically targeted optical electrophysiology in intact neural circuits. Cell. 2013;154:904–913. [Europe PMC free article] [Abstract] [Google Scholar]
81. St-Pierre F, Marshall JD, Yang Y, Gong Y, Schnitzer MJ, Lin MZ. High-fidelity optical reporting of neuronal electrical activity with an ultrafast fluorescent voltage sensor. Nat Neurosci. 2014;17:884–889. [Europe PMC free article] [Abstract] [Google Scholar]
82. Yang HH, St-Pierre F, Sun X, Ding X, Lin MZ, Clandinin TR. Subcellular imaging of voltage and calcium signals reveals neural processing in vivo. Cell. 2016;166:245–257. [Europe PMC free article] [Abstract] [Google Scholar]
83. Gong Y, Wagner MJ, Zhong Li J, Schnitzer MJ. Imaging neural spiking in brain tissue using FRET-opsin protein voltage sensors. Nat Commun. 2014;5:3674. [Europe PMC free article] [Abstract] [Google Scholar]
84. Aplin AC, Kaufman TC. Homeotic transformation of legs to mouthparts by proboscipedia expression in Drosophila imaginal discs. Mech Develop. 1997;62:51–60. [Abstract] [Google Scholar]
85. Zhou L, Schnitzler A, Agapite J, Schwartz LM, Steller H, Nambu JR. Cooperative functions of the reaper and head involution defective genes in the programmed cell death of Drosophila central nervous system midline cells. Proc Natl Acad Sci U S A. 1997;94:5131–5136. [Europe PMC free article] [Abstract] [Google Scholar]
86. Flood TF, Iguchi S, Gorczyca M, White B, Ito K, Yoshihara M. A single pair of interneurons commands the Drosophila feeding motor program. Nature. 2013;499:83–87. [Europe PMC free article] [Abstract] [Google Scholar]
87. Keller A, Sweeney ST, Zars T, O’Kane CJ, Heisenberg M. Targeted expression of tetanus neurotoxin interferes with behavioral responses to sensory input in Drosophila. J Neurobiol. 2002;50:221–233. [Abstract] [Google Scholar]
88. Kitamoto T. Conditional modification of behavior in Drosophila by targeted expression of a temperature-sensitive shibire allele in defined neurons. J Neurobiol. 2001;47:81–92. [Abstract] [Google Scholar]
89. Nitabach MN, Wu Y, Sheeba V, Lemon WC, Strumbos J, Zelensky PK, et al. Electrical hyperexcitation of lateral ventral pacemaker neurons desynchronizes downstream circadian oscillators in the fly circadian circuit and induces multiple behavioral periods. J Neurosci. 2006;26:479–489. [Europe PMC free article] [Abstract] [Google Scholar]
90. Baines RA, Uhler JP, Thompson A, Sweeney ST, Bate M. Altered electrical properties in Drosophila neurons developing without synaptic transmission. J Neurosci. 2001;21:1523–1531. [Europe PMC free article] [Abstract] [Google Scholar]
91. Nitabach MN, Blau J, Holmes TC. Electrical silencing of Drosophila pacemaker neurons stops the free-running circadian clock. Cell. 2002;109:485–495. [Abstract] [Google Scholar]
92. Datta SR, Vasconcelos ML, Ruta V, Luo S, Wong A, Demir E, et al. The Drosophila pheromone cVA activates a sexually dimorphic neural circuit. Nature. 2008;452:473–477. [Abstract] [Google Scholar]
93. Ruta V, Datta SR, Vasconcelos ML, Freeland J, Looger LL, Axel R. A dimorphic pheromone circuit in Drosophila from sensory input to descending output. Nature. 2010;468:686–690. [Abstract] [Google Scholar]
94. Rosenzweig M, Brennan KM, Tayler TD, Phelps PO, Patapoutian A, Garrity PA. The Drosophila ortholog of vertebrate TRPA1 regulates thermotaxis. Genes Dev. 2005;19:419–424. [Europe PMC free article] [Abstract] [Google Scholar]
95. Zhang W, Ge W, Wang Z. A toolbox for light control of Drosophila behaviors through Channelrhodopsin 2-mediated photoactivation of targeted neurons. Eur J Neurosci. 2007;26:2405–2416. [Abstract] [Google Scholar]
96. Inagaki HK, Jung Y, Hoopfer ED, Wong AM, Mishra N, Lin JY, et al. Optogenetic control of Drosophila using a red-shifted Channelrhodopsin reveals experience-dependent influences on courtship. Nat Methods. 2014;11:325–332. [Europe PMC free article] [Abstract] [Google Scholar]
97. Klapoetke NC, Murata Y, Kim SS, Pulver SR, Birdsey-Benson A, Cho YK, et al. Independent optical excitation of distinct neural populations. Nat Methods. 2014;11:338–346. [Europe PMC free article] [Abstract] [Google Scholar]
98. Zhang F, Wang LP, Brauner M, Liewald JF, Kay K, Watzke N, et al. Multimodal fast optical interrogation of neural circuitry. Nature. 2007;446:633–639. [Abstract] [Google Scholar]
99. Inada K, Kohsaka H, Takasu E, Matsunaga T, Nose A. Optical dissection of neural circuits responsible for Drosophila larval locomotion with Halorhodopsin. PLoS One. 2011;6:e29019. [Europe PMC free article] [Abstract] [Google Scholar]
100. Doll CA, Broadie K. Activity-dependent FMRP requirements in development of the neural circuitry of learning and memory. Development. 2015;142:1346–1356. [Europe PMC free article] [Abstract] [Google Scholar]
101. Mohammad F, Stewart JC, Ott S, Chlebikova K, Chua JY, Koh TW, et al. Optogenetic inhibition of behavior with anion Channelrhodopsins. Nat Methods. 2017;14:271–274. [Abstract] [Google Scholar]
102. Lima SQ, Miesenbock G. Remote control of behavior through genetically targeted photostimulation of neurons. Cell. 2005;121:141–152. [Abstract] [Google Scholar]
103. Patterson GH, Lippincott-Schwartz J. A photoactivatable GFP for selective photolabeling of proteins and cells. Science. 2002;297:1873–1877. [Abstract] [Google Scholar]
104. Sun Y, Nern A, Franconville R, Dana H, Schreiter ER, Looger LL, et al. Neural signatures of dynamic stimulus selection in Drosophila. Nat Neurosci. 2017;20:1104–1113. [Abstract] [Google Scholar]
105. Talay M, Richman EB, Snell NJ, Hartmann GG, Fisher JD, Sorkac A, et al. Transsynaptic mapping of second-order taste neurons in flies by trans-Tango. Neuron. 2017;96:783–795. [Europe PMC free article] [Abstract] [Google Scholar]
106. Lamaze A, Kratschmer P, Chen KF, Lowe S, Jepson JEC. A wake-promoting circadian output circuit in Drosophila. Curr Biol. 2018;28:3098–3105. [Abstract] [Google Scholar]
107. Huang TH, Niesman P, Arasu D, Lee D, De La Cruz AL, Callejas A, et al. Tracing neuronal circuits in transgenic animals by transneuronal control of transcription (TRACT) Elife. 2017;6:e32027. [Europe PMC free article] [Abstract] [Google Scholar]
108. Gordon MD, Scott K. Motor control in a Drosophila taste circuit. Neuron. 2009;61:373–384. [Europe PMC free article] [Abstract] [Google Scholar]
109. Fan P, Manoli DS, Ahmed OM, Chen Y, Agarwal N, Kwong S, et al. Genetic and neural mechanisms that inhibit Drosophila from mating with other species. Cell. 2013;154:89–102. [Europe PMC free article] [Abstract] [Google Scholar]
110. Macpherson LJ, Zaharieva EE, Kearney PJ, Alpert MH, Lin TY, Turan Z, et al. Dynamic labelling of neural connections in multiple colours by trans-synaptic fluorescence complementation. Nat Commun. 2015;6:10024. [Europe PMC free article] [Abstract] [Google Scholar]
111. Shearin HK, Quinn CD, Mackin RD, Macdonald IS, Stowers RS. t-GRASP, a targeted GRASP for assessing neuronal connectivity. J Neurosci Methods. 2018;306:94–102. [Europe PMC free article] [Abstract] [Google Scholar]
112. Chen Y, Akin O, Nern A, Tsui CY, Pecot MY, Zipursky SL. Cell-type-specific labeling of synapses in vivo through synaptic tagging with recombination. Neuron. 2014;81:280–293. [Europe PMC free article] [Abstract] [Google Scholar]
113. Ryan MD, Drew J. Foot-and-mouth disease virus 2A oligopeptide mediated cleavage of an artificial polyprotein. EMBO J. 1994;13:928–933. [Europe PMC free article] [Abstract] [Google Scholar]
114. Sugie A, Hakeda-Suzuki S, Suzuki E, Silies M, Shimozono M, Mohl C, et al. Molecular remodeling of the presynaptic active zone of Drosophila photoreceptors via activity-dependent feedback. Neuron. 2015;86:711–725. [Abstract] [Google Scholar]
115. Liu S, Liu Q, Tabuchi M, Wu MN. Sleep drive is encoded by neural plastic changes in a dedicated Circuit. Cell. 2016;165:1347–1360. [Europe PMC free article] [Abstract] [Google Scholar]
116. Zhang X, Li Q, Wang L, Liu ZJ, Zhong Y. Active protection: learning-activated Raf/MAPK activity protects labile memory from Rac1-independent forgetting. Neuron. 2018;98:142–155. [Abstract] [Google Scholar]
117. Shirangi TR, Wong AM, Truman JW, Stern DL. Doublesex regulates the connectivity of a neural circuit controlling Drosophila male courtship song. Dev Cell. 2016;37:533–544. [Abstract] [Google Scholar]
118. Felsenberg J, Barnstedt O, Cognigni P, Lin S, Waddell S. Re-evaluation of learned information in Drosophila. Nature. 2017;544:240–244. [Europe PMC free article] [Abstract] [Google Scholar]
119. Watanabe K, Chiu H, Pfeiffer BD, Wong AM, Hoopfer ED, Rubin GM, et al. A circuit node that integrates convergent input from neuromodulatory and social behavior-promoting neurons to control aggression in Drosophila. Neuron. 2017;95:1112–1128. [Europe PMC free article] [Abstract] [Google Scholar]
120. Green J, Adachi A, Shah KK, Hirokawa JD, Magani PS, Maimon G. A neural circuit architecture for angular integration in Drosophila. Nature. 2017;546:101–106. [Europe PMC free article] [Abstract] [Google Scholar]
121. Klapoetke NC, Nern A, Peek MY, Rogers EM, Breads P, Rubin GM, et al. Ultra-selective looming detection from radial motion opponency. Nature. 2017;551:237–241. [Europe PMC free article] [Abstract] [Google Scholar]
122. Chiang AS, Lin CY, Chuang CC, Chang HM, Hsieh CH, Yeh CW, et al. Three-dimensional reconstruction of brain-wide wiring networks in Drosophila at single-cell resolution. Curr Biol. 2011;21:1–11. [Abstract] [Google Scholar]
123. Shih CT, Sporns O, Yuan SL, Su TS, Lin YJ, Chuang CC, et al. Connectomics-based analysis of information flow in the Drosophila brain. Curr Biol. 2015;25:1249–1258. [Abstract] [Google Scholar]
124. Lin CY, Chuang CC, Hua TE, Chen CC, Dickson BJ, Greenspan RJ, et al. A comprehensive wiring diagram of the protocerebral bridge for visual information processing in the Drosophila brain. Cell Rep. 2013;3:1739–1753. [Abstract] [Google Scholar]
125. Costa M, Manton JD, Ostrovsky AD, Prohaska S, Jefferis GS. NBLAST: rapid, sensitive comparison of neuronal structure and construction of neuron family databases. Neuron. 2016;91:293–311. [Europe PMC free article] [Abstract] [Google Scholar]
126. Jeanne JM, Fisek M, Wilson RI. The organization of projections from olfactory glomeruli onto higher-order neurons. Neuron. 2018;98:1198–1213. [Europe PMC free article] [Abstract] [Google Scholar]
127. Vogelstein JT, Park Y, Ohyama T, Kerr RA, Truman JW, Priebe CE, et al. Discovery of brainwide neural-behavioral maps via multiscale unsupervised structure learning. Science. 2014;344:386–392. [Abstract] [Google Scholar]
128. Robie AA, Hirokawa J, Edwards AW, Umayam LA, Lee A, Phillips ML, et al. Mapping the neural substrates of behavior. Cell. 2017;170:393–406. [Abstract] [Google Scholar]
129. Briggman KL, Bock DD. Volume electron microscopy for neuronal circuit reconstruction. Curr Opin Neurobiol. 2012;22:154–161. [Abstract] [Google Scholar]
130. Takemura SY, Bharioke A, Lu Z, Nern A, Vitaladevuni S, Rivlin PK, et al. A visual motion detection circuit suggested by Drosophila connectomics. Nature. 2013;500:175–181. [Europe PMC free article] [Abstract] [Google Scholar]
131. Takemura SY. Connectome of the fly visual circuitry. Microscopy. 2015;64:37–44. [Abstract] [Google Scholar]
132. Ohyama T, Schneider-Mizell CM, Fetter RD, Aleman JV, Franconville R, Rivera-Alba M, et al. A multilevel multimodal circuit enhances action selection in Drosophila. Nature. 2015;520:633–639. [Abstract] [Google Scholar]
133. Berck ME, Khandelwal A, Claus L, Hernandez-Nunez L, Si G, Tabone CJ, et al. The wiring diagram of a glomerular olfactory system. Elife. 2016;5:e14859. [Europe PMC free article] [Abstract] [Google Scholar]
134. Takemura SY, Aso Y, Hige T, Wong A, Lu Z, Xu CS, et al. A connectome of a learning and memory center in the adult Drosophila brain. Elife. 2017;6:e26975. [Europe PMC free article] [Abstract] [Google Scholar]
135. Tobin WF, Wilson RI, Lee WA. Wiring variations that enable and constrain neural computation in a sensory microcircuit. Elife. 2017;6:e24838. [Europe PMC free article] [Abstract] [Google Scholar]
136. Zheng Z, Lauritzen JS, Perlman E, Robinson CG, Nichols M, Milkie D, et al. A complete electron microscopy volume of the brain of adult Drosophila melanogaster. Cell. 2018;174:730–743. [Europe PMC free article] [Abstract] [Google Scholar]
137. Bargmann CI. Beyond the connectome: how neuromodulators shape neural circuits. Bioessays. 2012;34:458–465. [Abstract] [Google Scholar]
138. Hassenstein B, Reichardt W. Systemtheoretische analyse der zeit-, reihenfolgen- und vorzeichenauswertung bei der bewegungsperzeption des Rüsselkäfers chlorophanus. Zeitschrift für Naturforschung B. 1956;11:513–524. [Google Scholar]
139. Clark DA, Bursztyn L, Horowitz MA, Schnitzer MJ, Clandinin TR. Defining the computational structure of the motion detector in Drosophila. Neuron. 2011;70:1165–1177. [Europe PMC free article] [Abstract] [Google Scholar]
140. Kim SS, Rouault H, Druckmann S, Jayaraman V. Ring attractor dynamics in the Drosophila central brain. Science. 2017;356:849–853. [Abstract] [Google Scholar]
141. Ito K, Shinomiya K, Ito M, Armstrong JD, Boyan G, Hartenstein V, et al. A systematic nomenclature for the insect brain. Neuron. 2014;81:755–765. [Abstract] [Google Scholar]
142. Stern DL, Crocker J, Ding Y, Frankel N, Kappes G, Kim E, et al. Genetic and transgenic reagents for Drosophila simulans, D. mauritiana, D. yakuba, D. santomea, and D. virilis. G3 (Bethesda) 2017, 7: 1339–1347. [Europe PMC free article] [Abstract]
143. Prieto-Godino LL, Rytz R, Cruchet S, Bargeton B, Abuin L, Silbering AF, et al. Evolution of acid-sensing olfactory circuits in drosophilids. Neuron. 2017;93:661–676. [Abstract] [Google Scholar]
144. Seeholzer LF, Seppo M, Stern DL, Ruta V. Evolution of a central neural circuit underlies Drosophila mate preferences. Nature. 2018;559:564–569. [Europe PMC free article] [Abstract] [Google Scholar]

Articles from Neuroscience Bulletin are provided here courtesy of Springer

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/61011045
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/61011045

Article citations


Go to all (18) article citations

Data 


Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.