Filter
Associated Lab
- Ahrens Lab (2) Apply Ahrens Lab filter
- Aso Lab (3) Apply Aso Lab filter
- Baker Lab (1) Apply Baker Lab filter
- Betzig Lab (8) Apply Betzig Lab filter
- Bock Lab (1) Apply Bock Lab filter
- Branson Lab (7) Apply Branson Lab filter
- Card Lab (4) Apply Card Lab filter
- Cardona Lab (8) Apply Cardona Lab filter
- Cui Lab (1) Apply Cui Lab filter
- Dickson Lab (1) Apply Dickson Lab filter
- Druckmann Lab (3) Apply Druckmann Lab filter
- Dudman Lab (4) Apply Dudman Lab filter
- Eddy/Rivas Lab (1) Apply Eddy/Rivas Lab filter
- Feliciano Lab (1) Apply Feliciano Lab filter
- Fetter Lab (4) Apply Fetter Lab filter
- Funke Lab (1) Apply Funke Lab filter
- Gonen Lab (11) Apply Gonen Lab filter
- Grigorieff Lab (6) Apply Grigorieff Lab filter
- Harris Lab (5) Apply Harris Lab filter
- Heberlein Lab (1) Apply Heberlein Lab filter
- Hermundstad Lab (1) Apply Hermundstad Lab filter
- Hess Lab (4) Apply Hess Lab filter
- Jayaraman Lab (4) Apply Jayaraman Lab filter
- Ji Lab (5) Apply Ji Lab filter
- Keleman Lab (1) Apply Keleman Lab filter
- Keller Lab (2) Apply Keller Lab filter
- Lavis Lab (16) Apply Lavis Lab filter
- Lee (Albert) Lab (6) Apply Lee (Albert) Lab filter
- Leonardo Lab (2) Apply Leonardo Lab filter
- Lippincott-Schwartz Lab (9) Apply Lippincott-Schwartz Lab filter
- Liu (Zhe) Lab (5) Apply Liu (Zhe) Lab filter
- Looger Lab (6) Apply Looger Lab filter
- Magee Lab (2) Apply Magee Lab filter
- Menon Lab (1) Apply Menon Lab filter
- Pachitariu Lab (1) Apply Pachitariu Lab filter
- Reiser Lab (6) Apply Reiser Lab filter
- Riddiford Lab (1) Apply Riddiford Lab filter
- Romani Lab (6) Apply Romani Lab filter
- Rubin Lab (15) Apply Rubin Lab filter
- Saalfeld Lab (4) Apply Saalfeld Lab filter
- Scheffer Lab (4) Apply Scheffer Lab filter
- Schreiter Lab (4) Apply Schreiter Lab filter
- Shroff Lab (1) Apply Shroff Lab filter
- Simpson Lab (2) Apply Simpson Lab filter
- Singer Lab (6) Apply Singer Lab filter
- Spruston Lab (1) Apply Spruston Lab filter
- Stern Lab (8) Apply Stern Lab filter
- Sternson Lab (2) Apply Sternson Lab filter
- Svoboda Lab (9) Apply Svoboda Lab filter
- Truman Lab (6) Apply Truman Lab filter
- Turaga Lab (3) Apply Turaga Lab filter
- Turner Lab (2) Apply Turner Lab filter
- Wu Lab (1) Apply Wu Lab filter
- Zlatic Lab (7) Apply Zlatic Lab filter
Associated Project Team
- Fly Descending Interneuron (1) Apply Fly Descending Interneuron filter
- Fly Functional Connectome (4) Apply Fly Functional Connectome filter
- Fly Olympiad (1) Apply Fly Olympiad filter
- FlyEM (4) Apply FlyEM filter
- FlyLight (2) Apply FlyLight filter
- GENIE (3) Apply GENIE filter
- ThalamoSeq (1) Apply ThalamoSeq filter
- Tool Translation Team (T3) (3) Apply Tool Translation Team (T3) filter
- Transcription Imaging (6) Apply Transcription Imaging filter
Associated Support Team
- Anatomy and Histology (2) Apply Anatomy and Histology filter
- Cryo-Electron Microscopy (4) Apply Cryo-Electron Microscopy filter
- Electron Microscopy (1) Apply Electron Microscopy filter
- Integrative Imaging (1) Apply Integrative Imaging filter
- Invertebrate Shared Resource (1) Apply Invertebrate Shared Resource filter
- Project Technical Resources (1) Apply Project Technical Resources filter
- Quantitative Genomics (2) Apply Quantitative Genomics filter
- Scientific Computing Software (9) Apply Scientific Computing Software filter
- Viral Tools (1) Apply Viral Tools filter
- Vivarium (1) Apply Vivarium filter
Publication Date
- December 2017 (15) Apply December 2017 filter
- November 2017 (11) Apply November 2017 filter
- October 2017 (7) Apply October 2017 filter
- September 2017 (14) Apply September 2017 filter
- August 2017 (15) Apply August 2017 filter
- July 2017 (20) Apply July 2017 filter
- June 2017 (18) Apply June 2017 filter
- May 2017 (25) Apply May 2017 filter
- April 2017 (21) Apply April 2017 filter
- March 2017 (15) Apply March 2017 filter
- February 2017 (7) Apply February 2017 filter
- January 2017 (18) Apply January 2017 filter
- Remove 2017 filter 2017
186 Janelia Publications
Showing 151-160 of 186 resultsArfA rescues ribosomes stalled on truncated mRNAs by recruiting release factor RF2, which normally binds stop codons to catalyze peptide release. We report two 3.2-Å resolution cryo-EM structures - determined from a single sample - of the 70S ribosome with ArfA•RF2 in the A site. In both states, the ArfA C-terminus occupies the mRNA tunnel downstream of the A site. One state contains a compact inactive RF2 conformation. Ordering of the ArfA N-terminus in the second state rearranges RF2 into an extended conformation that docks the catalytic GGQ motif into the peptidyl-transferase center. Our work thus reveals the structural dynamics of ribosome rescue. The structures demonstrate how ArfA "senses" the vacant mRNA tunnel and activates RF2 to mediate peptide release without a stop codon, allowing stalled ribosomes to be recycled.
We recently identified ten novel SLE susceptibility loci in Asians and uncovered several additional suggestive loci requiring further validation. This study aimed to replicate five of these suggestive loci in a Han Chinese cohort from Hong Kong, followed by meta-analysis (11,656 cases and 23,968 controls) on previously reported Asian and European populations, and to perform bioinformatic analyses on all 82 reported SLE loci to identify shared regulatory signatures. We performed a battery of analyses for these five loci, as well as joint analyses on all 82 SLE loci. All five loci passed genome-wide significance: MYNN (rs10936599, Pmeta = 1.92 × 10-13, OR = 1.14), ATG16L2 (rs11235604, Pmeta = 8.87 × 10 -12, OR = 0.78), CCL22 (rs223881, Pmeta = 5.87 × 10-16, OR = 0.87), ANKS1A (rs2762340, Pmeta = 4.93 × 10-15, OR = 0.87) and RNASEH2C (rs1308020, Pmeta = 2.96 × 10-19, OR = 0.84) and co-located with annotated gene regulatory elements. The novel loci share genetic signatures with other reported SLE loci, including effects on gene expression, transcription factor binding, and epigenetic characteristics. Most (56%) of the correlated (r2 > 0.8) SNPs from the 82 SLE loci were implicated in differential expression (9.81 × 10-198 < P < 5 × 10-3) of cis-genes. Transcription factor binding sites for p53, MEF2A and E2F1 were significantly (P < 0.05) over-represented in SLE loci, consistent with apoptosis playing a critical role in SLE. Enrichment analysis revealed common pathways, gene ontology, protein domains, and cell type-specific expression. In summary, we provide evidence of five novel SLE susceptibility loci. Integrated bioinformatics using all 82 loci revealed that SLE susceptibility loci share many gene regulatory features, suggestive of conserved mechanisms of SLE etiopathogenesis.
Progressive, technological achievements in the quantitative fluorescence microscopy field are allowing researches from many different areas to start unraveling the dynamic intricacies of biological processes inside living cells. From super-resolution microscopy techniques to tracking of individual proteins, fluorescence microscopy is changing our perspective on how the cell works. Fortunately, a growing number of research groups are exploring single-molecule studies in living cells. However, no clear consensus exists on several key aspects of the technique such as image acquisition conditions, or analysis of the obtained data. Here, we describe a detailed approach to perform single-molecule tracking (SMT) of transcription factors in living cells to obtain key binding characteristics, namely their residence time and bound fractions. We discuss different types of fluorophores, labeling density, microscope, cameras, data acquisition, and data analysis. Using the glucocorticoid receptor as a model transcription factor, we compared alternate tags (GFP, mEOS, HaloTag, SNAP-tag, CLIP-tag) for potential multicolor applications. We also examine different methods to extract the dissociation rates and compare them with simulated data. Finally, we discuss several challenges that this exciting technique still faces.
Our ability to unambiguously image and track individual molecules in live cells is limited by packing of multiple copies of labeled molecules within the resolution limit. Here we devise a universal genetic strategy to precisely control protein copy number in a cell. This system has a dynamic titration range of more than 10,000 fold, enabling sparse labeling of proteins expressed at widely different levels. Combined with fluorescence signal amplification tags, this system extends the duration of automated single-molecule tracking by 2 orders of magnitude. We demonstrate long-term imaging of synaptic vesicle dynamics in cultured neurons as well as in live zebrafish. We found that axon initial segment utilizes a waterfall mechanism gating synaptic vesicle transport polarity by promoting anterograde transport processivity. Long-time observation also reveals that transcription factor Sox2 samples clustered binding sites in spatially-restricted sub-nuclear regions, suggesting that topological structures in the nucleus shape local gene activities by a sequestering mechanism. This strategy thus greatly expands the spatiotemporal length scales of live-cell single-molecule measurements for a quantitative understanding of complex control of molecular dynamics in vivo.
View Publication PageSpecies of the Drosophila melanogaster species subgroup, including the species D. simulans, D. mauritiana, D. yakuba, and D. santomea, have long served as model systems for studying evolution. Studies in these species have been limited, however, by a paucity of genetic and transgenic reagents. Here we describe a collection of transgenic and genetic strains generated to facilitate genetic studies within and between these species. We have generated many strains of each species containing mapped piggyBac transposons including an enhanced yellow fluorescent protein gene expressed in the eyes and a phiC31 attP site-specific integration site. We have tested a subset of these lines for integration efficiency and reporter gene expression levels. We have also generated a smaller collection of other lines expressing other genetically encoded fluorescent molecules in the eyes and a number of other transgenic reagents that will be useful for functional studies in these species. In addition, we have mapped the insertion locations of 58 transposable elements in D. virilis that will be useful for genetic mapping studies.
RNA-protein interactions are essential for proper gene expression regulation, particularly in neurons with unique spatial constraints. Currently, these interactions are defined biochemically, but a method is needed to evaluate them quantitatively within morphological context. Colocalization of two-color labels using wide-field microscopy is a method to infer these interactions. However, because of chromatic aberrations in the objective lens, this approach lacks the resolution to determine whether two molecules are physically in contact or simply nearby by chance. Here, we developed a robust super registration methodology that corrected the chromatic aberration across the entire image field to within 10 nm, which is capable of determining whether two molecules are physically interacting or simply in proximity by random chance. We applied this approach to image single-molecule FISH in combination with immunofluorescence (smFISH-IF) and determined whether the association between an mRNA and binding protein(s) within a neuron was significant or accidental. We evaluated several mRNA-binding proteins identified from RNA pulldown assays to determine which of these exhibit bona fide interactions. Surprisingly, many known mRNA-binding proteins did not bind the mRNA in situ, indicating that adventitious interactions are significant using existing technology. This method provides an ability to evaluate two-color registration compatible with the scale of molecular interactions.
Insects, like most animals, tend to steer away from imminent threats [1-7]. Drosophila melanogaster, for example, generally initiate an escape take-off in response to a looming visual stimulus, mimicking a potential predator [8]. The escape response to a visual threat is, however, flexible [9-12] and can alternatively consist of walking backward away from the perceived threat [11], which may be a more effective response to ambush predators such as nymphal praying mantids [7]. Flexibility in escape behavior may also add an element of unpredictability that makes it difficult for predators to anticipate or learn the prey's likely response [3-6]. Whereas the fly's escape jump has been well studied [8, 9, 13-18], the neuronal underpinnings of evasive walking remain largely unexplored. We previously reported the identification of a cluster of descending neurons-the moonwalker descending neurons (MDNs)-the activity of which is necessary and sufficient to trigger backward walking [19], as well as a population of visual projection neurons-the lobula columnar 16 (LC16) cells-that respond to looming visual stimuli and elicit backward walking and turning [11]. Given the similarity of their activation phenotypes, we hypothesized that LC16 neurons induce backward walking via MDNs and that turning while walking backward might reflect asymmetric activation of the left and right MDNs. Here, we present data from functional imaging, behavioral epistasis, and unilateral activation experiments that support these hypotheses. We conclude that LC16 and MDNs are critical components of the neural circuit that transduces threatening visual stimuli into directional locomotor output.
Endothelial exocytosis of Weibel-Palade body (WPB) is one of the first lines of defence against vascular injury. However, the mechanisms that control WPB exocytosis in the final stages (including the docking, priming and fusion of granules) are poorly understood. Here we show that the focal adhesion protein zyxin is crucial in this process. Zyxin downregulation inhibits the secretion of von Willebrand factor (VWF), the most abundant cargo in WPBs, from human primary endothelial cells (ECs) induced by cAMP agonists. Zyxin-deficient mice exhibit impaired epinephrine-stimulated VWF release, prolonged bleeding time and thrombosis, largely due to defective endothelial secretion of VWF. Using live-cell super-resolution microscopy, we visualize previously unappreciated reorganization of pre-existing actin filaments around WPBs before fusion, dependent on zyxin and an interaction with the actin crosslinker α-actinin. Our findings identify zyxin as a physiological regulator of endothelial exocytosis through reorganizing local actin network in the final stage of exocytosis.
Dopamine (DA) is a neurotransmitter with conserved behavioral roles between invertebrate and vertebrate animals. In addition to its neural functions, in insects DA is a critical substrate for cuticle pigmentation and hardening. Drosophila tyrosine hydroxylase (DTH) is the rate limiting enzyme for DA biosynthesis. Viable brain DA deficient flies were previously generated using tissue selective GAL4-UAS binary expression rescue of a DTH null mutation and these flies show specific behavioral impairments. To circumvent the limitations of rescue via binary expression, here we achieve rescue utilizing genomically integrated mutant DTH. As expected, our DA deficient flies have no detectable DTH or DA in the brain, and show reduced locomotor activity. This deficit can be rescued by L-DOPA/carbidopa feeding, similar to human Parkinson's disease treatment. Genetic rescue via GAL4/UAS-DTH was also successful, although this required the generation of a new UAS-DTH1 transgene devoid of most untranslated regions, since existing UAS-DTH transgenes express in the brain without a Gal4 driver via endogenous regulatory elements. A surprising finding of our newly constructed UAS-DTH1m is that it expresses DTH at an undetectable level when regulated by dopaminergic GAL4 drivers even when fully rescuing DA, indicating that DTH immunostaining is not necessarily a valid marker for DA expression. This finding necessitated optimizing DA immunohistochemistry, revealing details of DA innervation to the mushroom body and the central complex. When DA rescue is limited to specific DA neurons, DA does not diffuse beyond the DTH-expressing terminals, such that DA signaling can be limited to very specific brain regions.
The calcium-modulated photoactivatable ratiometric integrator CaMPARI (Fosque et al., 2015) facilitates the study of neural circuits by permanently marking cells active during user-specified temporal windows. Permanent marking enables measurement of signals from large swathes of tissue and easy correlation of activity with other structural or functional labels. One potential application of CaMPARI is labeling neurons postsynaptic to specific populations targeted for optogenetic stimulation, giving rise to all-optical functional connectivity mapping. Here, we characterized the response of CaMPARI to several common types of neuronal calcium signals in mouse acute cortical brain slices. Our experiments show that CaMPARI is effectively converted by both action potentials and sub-threshold synaptic inputs, and that conversion level is correlated to synaptic strength. Importantly, we found that conversion rate can be tuned: it is linearly related to light intensity. At low photoconversion light levels CaMPARI offers a wide dynamic range due to slower conversion rate; at high light levels conversion is more rapid and more sensitive to activity. Finally, we employed CaMPARI and optogenetics for functional circuit mapping in ex vivo acute brain slices, which preserve in vivo-like connectivity of axon terminals. With a single light source, we stimulated channelrhodopsin-2-expressing long-range posteromedial (POm) thalamic axon terminals in cortex and induced CaMPARI conversion in recipient cortical neurons. We found that POm stimulation triggers robust photoconversion of layer 5 cortical neurons and weaker conversion of layer 2/3 neurons. Thus, CaMPARI enables network-wide, tunable, all-optical functional circuit mapping that captures supra- and sub-threshold depolarization. This article is protected by copyright. All rights reserved.