Skip to main content
Advertisement
Main content starts here

Taking a snapshot of active brain circuitry

Neuroscientists now have a method to mark active populations of neurons in vivo to study circuit activity in the behaving animal. Fosque et al. designed and thoroughly validated a fluorescent protein–based reagent that allows permanent marking of active cells over short time scales. This indicator, termed CaMPARI, switches from its native green to a red fluorescent state by simultaneous illumination with violet light and exposure to increased levels of intracellular calcium. CaMPARI successfully marked active nerve cells in Drosophila, zebrafish, and mouse brains.
Science, this issue p. 755

Abstract

The identification of active neurons and circuits in vivo is a fundamental challenge in understanding the neural basis of behavior. Genetically encoded calcium (Ca2+) indicators (GECIs) enable quantitative monitoring of cellular-resolution activity during behavior. However, such indicators require online monitoring within a limited field of view. Alternatively, post hoc staining of immediate early genes (IEGs) indicates highly active cells within the entire brain, albeit with poor temporal resolution. We designed a fluorescent sensor, CaMPARI, that combines the genetic targetability and quantitative link to neural activity of GECIs with the permanent, large-scale labeling of IEGs, allowing a temporally precise “activity snapshot” of a large tissue volume. CaMPARI undergoes efficient and irreversible green-to-red conversion only when elevated intracellular Ca2+ and experimenter-controlled illumination coincide. We demonstrate the utility of CaMPARI in freely moving larvae of zebrafish and flies, and in head-fixed mice and adult flies.
Brain function relies on patterns of synaptic input and action potential firing, which are accompanied by transient changes in free intracellular calcium (Ca2+) concentration ([Ca2+]) (1, 2). Genetically encoded calcium indicators (GECIs) are useful for monitoring the activity of populations of neurons and synapses in behaving organisms (36). However, the transient nature of GECI responses following [Ca2+] rises requires continuous monitoring during behavior using sophisticated imaging equipment with limited fields of view, and often physical restraint to prevent brain movement. Alternatively, expression of immediate early genes (IEGs) such as Arc and cFos can be measured following free behavior (7) within several hours, but is only weakly correlated with neural electrical activity (8, 9) and is constrained neither by genetic cell type nor by a precise temporal window. The creation of molecular tools to allow genetic targeting of direct reporters of neural activity that can be rendered permanent for post hoc analysis across entire brains, and that only mark neurons active during short, user-defined behavioral epochs, would be transformative (10, 11).
The GCaMP calcium indicator reports changes in free [Ca2+] [rise and decay times <1 s (12)] through mechanisms that modulate the environment of the 488-nm–absorbing, fluorescent 4-(p-hydroxybenzylidene)-5-imidazolinone chromophore of circularly permuted green fluorescent protein (cpGFP) (13). The photoconvertible fluorescent protein (FP) EosFP is a bright green FP that irreversibly converts to a bright red fluorescent species {backbone cleavage produces 2-[(1E)-2-(5-imidazolyl)ethenyl]-4-(p-hydroxybenzylidene)-5-imidazolinone} upon illumination with violet light (14). Previously, circular permutation of a photoconvertible FP and attachment to the Ca2+-binding protein calmodulin (CaM) and its associated M13 peptide (M13) gave rise to a photoconvertible GECI that produced GCaMP-like rises in fluorescent intensity both in the unconverted, green state and in the converted, red state (15). We reasoned that related constructs would undergo Ca2+-dependent allosteric modulation of the chromophore photoconvertibility. Such a protein species would convert from green to red only in the simultaneous presence of high [Ca2+] and user-supplied violet light (Fig. 1A). By combining library screening and structure-guided mutagenesis (16) we produced such a protein, which we named CaMPARI (calcium-modulated photoactivatable ratiometric integrator).
Fig. 1 CaMPARI engineering and in vitro characterization.
(A) Schematic of CaMPARI function. (B) Primary structure of CaMPARI with mutations relative to mEos2. (C) Time course of red fluorescence appearance during exposure of CaMPARI to PC light. Lines represent single exponential fits to the PC time course. n = 3 measurements. All panels show means ± SEM. (D) CaMPARI PC rate as a function of free Ca2+. Black line represents a sigmoidal fit. n =3 measurements. (E) Relative fluorescence of CaMPARI as a function of free Ca2+ concentration. Lines are sigmoidal fits to the data. n =3 measurements. (F and G) Fluorescence of primary rat hippocampal neurons expressing CaMPARI after a 2-s PC light pulse without and with 80-Hz field electrode stimulation. (H) Color scale for composite micrographs of green and red CaMPARI fluorescence used throughout this work. (I) Quantitation of the red/green fluorescence ratio of neurons stimulated at different frequencies during PC light. Black line is a linear fit to the data. n = 10 measurements. (J) Fluorescence from two fixed, immunostained neurons. (K) Individual color channels from the same field of view as in (J). (L) Correlation between intensity of phospho-CREB staining and CaMPARI PC for high-potassium depolarization (90 mM K+), controls (4 mM K+), and P2X/ATP. A range of expression levels of P2X channels was present across neurons, resulting in the observed spread in activation upon 100 μM ATP stimulation. Scale bars: 50 μm in all panels.
CaMPARI has the primary structure depicted in Fig. 1B and fig. S1, and includes a nuclear export signal (NES) to exclude it from the nucleus when expressed in eukaryotic cells. CaMPARI exhibits 60 and 40% of the green and red fluorescence brightness of mEos2, respectively, and photoconverts to the red form 21 times as fast in the presence of calcium (Fig. 1C and table S1). Photoconversion (PC) of the calcium-bound and calcium-free states of CaMPARI occurs three times and one-seventh as fast, respectively, as that of the parent EosFP variant (fig. S2). No effect of calcium on PC of the parent EosFP variant was observed (fig. S2).
Measuring the PC rate while titrating calcium gave an apparent dissociation constant (Kd) of 128 ± 6 nM (Fig. 1D). We modified the affinity of CaMPARI for calcium via mutagenesis at the CaM/M13 interface (fig. S3), thus increasing dynamic range in cells with a wide range of baseline and peak calcium concentrations. In addition to a calcium-dependent green-to-red PC rate, the fluorescence of both green and red CaMPARI states is also calcium-dependent, decreasing by a factor of 9 and 23, respectively, upon calcium binding (Fig. 1E, fig. S4, and table S1).
After initial characterization in cultured HeLa cells (fig. S5), we expressed CaMPARI in primary cultured rat hippocampal neurons. Whereas a 2-s pulse of PC light (1.5 W/cm2) resulted in only minimal CaMPARI PC in unstimulated neurons, the same light exposure induced significant PC in field-stimulated neurons (Fig. 1, F to H), with a roughly linear correspondence between the frequency of field stimulation and the extent of CaMPARI PC (Fig. 1I). Stimulation of CaMPARI-expressing neurons without exposure to 405-nm light led to transient decreases of its green fluorescence (fig. S6), but with no conversion to the red form. We additionally depolarized neurons either by increasing extracellular potassium concentration or via pharmacogenetic activation of exogenously expressed P2X channels with adenosine 5′-triphosphate (ATP), both in concert with application of PC light. Cells were fixed with 4% paraformaldehyde (PFA) and colabeled with traditional antibody-staining techniques (Fig. 1, J and K). The green and red fluorescence of CaMPARI each decreased ~50% upon PFA fixation (fig. S7). Retention of the CaMPARI PC signal through fixation, permeabilization, and multicolor antibody labeling enables integrated Ca2+ concentrations within individual cells to be directly correlated to independent proteomic metrics typically used for post hoc estimation of cell-type identification and examination of a wide range of biochemical events related to plasticity and other key signaling cascades. We coimmunolabeled fixed neurons for P2X expression level (fig. S8, an indication of the magnitude of the exogenous pharmacogenetic manipulation) and phosphorylated nuclear cyclic adenosine 3′,5′-monophosphate (cAMP)–responsive element-binding protein (pCREB), which increased with elevated intracellular calcium (Fig. 1L). CaMPARI signal after fixation was similar to that of conventional immunofluorescence and in situ hybridization markers, making it possible to image a direct measure of Ca2+ in parallel with proteomic markers that can only be accessed after fixation.
We next used CaMPARI to mark active neurons in vivo, during sensory stimuli or behavior, in four preparations of three model organisms: larval zebrafish and Drosophila melanogaster to demonstrate whole-brain and genetically targeted neural ensemble mapping during free behavior; mouse primary visual cortex (V1) for large-scale post hoc confirmation of known distributions of direction-selective cells in a mammalian brain; and adult Drosophila to highlight “labeled lines” of synaptically connected ensembles following naturalistic or optogenetic stimulation of sensory pathways. We chose calcium affinity variants (CaMPARI in mouse and adult Drosophila; V398D, with about one-seventh the affinity for Ca2+, in larval and adult Drosophila; W391F+V398L, with about one-half the affinity for Ca2+, in larval zebrafish; fig. S3) based on performance of different GCaMP variants in these preparations, as well as literature estimates of intracellular calcium.
To demonstrate the utility of CaMPARI for whole-brain neural circuit marking in larval zebrafish (Danio rerio) during free behavior, we generated stable transgenic zebrafish expressing CaMPARI in all neurons from the elavl3 promoter (16) (Fig. 2A). After exposure of larvae (4 to 5 days after fertilization) to 10 s of PC light (405-nm light-emitting diode array, 400 mW/cm2, Fig. 2B), confocal stacks were acquired to generate a cellular-resolution snapshot of the calcium state of all neurons in the zebrafish brain (Fig. 2C and movie S1). Exposure of larvae anesthetized with tricaine methanesulfonate (MS-222) to PC light resulted in a lack of PC throughout the brain (Fig. 2C). Freely swimming fish with no additional stimulation showed a large amount of PC in the forebrain and habenula, but very little in the optic tectum. Additionally, patterns of hindbrain activity consistent with motor output for swimming (17, 18) were often observed. Treatment of larvae with proconvulsive compounds such as 4-aminopyridine (4-AP), exposure to noxious heat or cold stimuli (Fig. 2C), or exposure to turbulent water (fig. S9) during the PC light pulse resulted in qualitatively different patterns of CaMPARI PC throughout the brain, consistent with permanent marking of the subset of neurons activated by these stimuli. CaMPARI signals following each stimulus were consistent between fish (figs. S9 and S10) and showed clear cellular resolution (Fig. 2C, bottom panels).
Fig. 2 Response in freely moving organisms.
(A) CaMPARI expression in the Tg[elavl3:CaMPARI(W391F+V398L)]jf9 zebrafish. Dashed black rectangle represents the region displayed in the top panels of (C). (B) Schematic of experimental configuration. (C) Confocal images of zebrafish larvae (4 to 5 days after fertilization) after 10 s of PC light, applied during exposure to different conditions (labels above: MS-222 is tricaine, a sodium channel blocker; 4-AP is 4-aminopyridine, a potassium channel blocker; noxious heat is exposure to 45°C water; noxious cold is exposure to 4°C water). Top panels are maximum-intensity Z projections (MIPs) of the entire brain. Dashed gray box in the top left panel represents the region displayed in the bottom panels. Bottom panels are individual Z slices from the same fish, illustrating individual cells and neuropils within portions of the optic tectum (OT), cerebellum (CB), and hindbrain (HB). (D) Schematic of labeled peripheral sensory neurons (PSNs) in a Drosophila larva. Axons (green lines) project from the body wall to the brain and branch along tracts within the VNC. Dashed gray box represents the region displayed in (E). (E) Confocal images of CaMPARI fluorescence in a subset of neurons in the VNC, driven by P0163-Gal4. Rectangles are segmented axonal tracts. P, proprioceptive; C, chordotonal; and N, nociceptive. (F) Schematic of the experimental setup. (G) Quantitation of the red/green CaMPARI fluorescence in sensory neuron projections. Mean ± SEM from n = 9 larvae in each condition (**P = 0.003; n.s., not significant, P > 0.4; Student’s t tests). (H) Dose response in chordotonal axons. Mean ± SEM from n = 6 to 10 larvae in each condition. Scale bars: 500 μm in (A), 100 μm in (C), and 20 μm in (E).
Next, we tracked activity in a freely moving organism with CaMPARI expression isolated to a genetically specified subset of neurons. We expressed CaMPARI in peripheral sensory neurons (PSNs) of Drosophila melanogaster larvae using the P0163-Gal4 line (19). In third-instar larvae, CaMPARI fluorescence was clearly visible in PSN axonal projections in the ventral nerve cord (VNC) and could be segmented into proprioceptive, chordotonal, and nociceptive axon terminals (Fig. 2E). To highlight PSN responses to vibrational stimuli (20), we administered a single 5-s pulse of PC light to freely crawling larvae on an agar plate during presentation of a 1000-Hz tone delivered from a speaker below the plate (Fig. 2F). The vibration stimulus produced a twofold increase in green-to-red PC only within the axonal projections of chordotonal neurons, not the axons of proprioceptive or nociceptive neurons (Fig. 2G). The extent of CaMPARI PC within chordotonal axons increased in a dose-dependent manner with increasing amplitude of the vibration stimulus, with more sensitive responses in lateral compared with ventromedial axon terminals (Fig. 2H).
Within layer 2/3 of mouse V1, neurons responsive to different directions and orientations of moving bar gratings are interspersed throughout the tissue (21). CaMPARI was delivered to layer 2/3 pyramidal cells with an adeno-associated virus 1 (AAV1)-synapsin1 virus. Expression was bright and excluded from the nucleus (Fig. 3B). First, in vivo calcium imaging was performed through a cranial window by imaging two-photon excited CaMPARI fluorescence as moving gratings were presented to the contralateral eye (16) (Fig. 3A). This allowed generation of orientation-tuning maps of segmented cell bodies. After identification of cells responsive to specific orientations (Fig. 3C), the visual stimulus (a single direction— “northwest,” preferred by cell 1 but not cell 2 in Fig. 3B) was replayed, and PC light was delivered through the cranial window in 500-ms pulses. After 20 PC pulses spaced 12 s apart, noticeable red fluorescence was present in a subset of cells (cell 1, but not cell 2, in Fig. 3B and movie S2). Neurons responsive to the “northwest” moving grating direction (like cell 1 in Fig. 3) exhibited a red/green fluorescence ratio significantly higher than cells not responsive to that direction (like cell 2 in Fig. 3, fig. S11). We relocated the in vivo two-photon fields of view within fixed, sectioned tissue using a confocal microscope (Fig. 3, B and E, and fig. S12). The relative red/green ratio of cells was maintained after fixation (fig. S13), and we were additionally able to image the CaMPARI signal over a much larger volume (Fig. 3F) than would have been easily accessible through the cranial window in vivo. When perfusion and fixation occurred 24 hours after PC, marked cells could still be easily distinguished in fixed sections with confocal microscopy (fig. S14).
Fig. 3 Response in mouse primary visual cortex.
(A) Schematic of experimental setup. (B) Two-photon fluorescence from cortical layer 2/3 of V1 after visual stimulus and PC light pulses. Two cells are circled and labeled for reference. (C) Calcium imaging fluorescence traces of the same two cells in response to different directions of drifting gratings (arrows and lines above traces). Average baseline-normalized green CaMPARI fluorescence (F/F0) from five trials is in green; individual trial traces are light gray. Vertical gray bars represent periods of drifting grating visual stimulus. Vertical dashed lines show the timing of 500-ms PC light pulses during the “northwest” stimulus. (D) CaMPARI Fred/Fgreen of cells according to functional properties identified by calcium imaging. “PC tuned”: cells with a significant (analysis of variance test, P < 0.01) calcium imaging response component to “northwest” gratings, (n = 15); “responsive, not PC tuned”: cells with a significant response to at least one direction, but no response to “northwest” (n = 39); “non-responsive”: no significant response during any grating orientation (n = 248) (***P < 0.001; n.s., not significant; P = 0.13; Mann-Whitney U tests). (E) Confocal image of fixed, sectioned tissue showing the same field of view as in (B). (F) Stitched composite confocal image of fixed, sectioned tissue. The region shown in (E) is outlined with a dashed gray box. Scale bars: 20 μm in (B) and (E), 200 μm in (F).
Genetic tools in Drosophila do not include methods for trans-synaptic circuit tracing and whole-brain functional mapping. Although photoactivatable GFP (22) and optogenetics in combination with calcium imaging (23) have been employed to detect potential neuronal connections, these methods require knowledge of presynaptic neurons and operate on a limited field of view. We set out to test whether CaMPARI could provide a more comprehensive solution. We affixed intact flies with pan-neuronal expression of low-affinity CaMPARI (V398D) to a physiology holder via the head cuticle and delivered PC light through a water-dipping objective while stimulating the animals with a panel of odors (Fig. 4A and fig. S15): (i) 3-octanol (3-Oct), which activates multiple glomeruli in the antennal lobe [AL; see (16) for a list of anatomical abbreviations used] (24); (ii) geosmin (Geo), an odorant that specifically activates the DA2 glomerulus (25); and (iii) phenylacetic acid (PAA), which strongly activates ionotropic receptor neurons projecting to the VL2a glomerulus (26). CaMPARI mapping confirmed the known neural representations of these odors in the antennal lobe (Fig. 4C, figs. S16 and S17, and movies S3 to S8). We also found additional, previously undescribed representations of some odors (Fig. 4, D and E). In the previous Geo study (25), a specific driver, GH146-Gal4, was used to express GCaMP3 in a limited subset of cells, facilitating segmentation of live-cell imaging data at the cost of ensemble undersampling. We similarly saw only the DA2 glomerulus activated when we drove CaMPARI expression with GH146-Gal4 (fig. S18), but additionally observed the DC1, VA1d, VA1v, and VM7 glomeruli (Fig. 4, D and E) with pan-neuronal CaMPARI. In the case of PAA, the original study (26) assessed odor responses using electrophysiology from specific Ir84a receptor neurons projecting to a single glomerulus (VL2a), rather than sampling the entire AL. When pan-neuronal CaMPARI was used, VL2a was the most salient responder, but weaker PC of VA1v and VM7 was also seen. Increasing the stimulation and PC time increased the overall red/green ratio without altering the relative amount of signal in different glomeruli (Fig. 4F).
Fig. 4 Functional mapping and circuit tracing in adult Drosophila.
(A) Schematic of experimental configuration. PC light is directed through an objective onto the head cuticle of a head-fixed adult fly during odor (C to F) or optogenetic (G to L) stimulation. (B) Schematic of olfactory circuitry targeted. The central brain of an adult fly is represented as a gray outline. ORNs [see (16) for a list of anatomical abbreviations used] (green) expressing specific receptors converge onto cognate glomeruli in the AL. PNs (red) from these glomeruli target the CA and LH. Putative tertiary neurons (cyan) then project to several areas including VL/SIP, SLP, VLP and finally to DN (orange), into the VNC. Dashed black box represents the area depicted in (C), (D), and (E). (C) MIPs of CaMPARI fluorescence from the left AL in response to indicated odors during PC. (D) MIP of red/green fluorescence ratios from (C). Dotted lines outline responsive glomeruli. (E) Maps of odor-response patterns in the AL. Each colored shape outlines one glomerulus from (C) and (D). Filled red glomeruli were previously reported in literature as responsive to that odor. Glomeruli that could be identified are labeled. (F) Comparison of glomerulus-specific responses to PAA at different PC and stimulation durations (100 s, n = 4; 200 s, n = 4; 300 s, n = 3). Means ± SEM. (G to L) CsChrimson is expressed in specific ORNs while CaMPARI is expressed pan-neuronally. (G) Schematic showing the proposed connectivity of the olfactory circuit analyzed. (H) Response pattern in AL shown with MIP; Ir84a ORNs are colabeled with mVenus. Ir84a-LexA.VP16 drives ORNs innervating ventromedial and dorsomedial glomeruli (see supplementary materials) that are photoconverted. (I) MIP of posterior sections showing marked PN axons targeting CA via the mALT bundle and LH through both the mALT and mlALT bundles. Also shown are labeled putative tertiary neurons in VLP. (J) MIP of more medial sections [relative to (I), same brain] showing the mlALT tract of PNs marked. (K and L) Experiments using high-affinity CaMPARI. (K) Similar to (H). (L) MIP of more medial sections [relative to (K), same brain] showing marked primary (ORNs, green arrows), secondary (PNs, red arrow), and putative tertiary (cyan arrows) neurons after optogenetic activation–PC. Scale bars: 25 μm in (C) and 50 μm in (H) to (L).
Olfactory receptor neuron (ORN) responses likely contribute substantially to the activation patterns we observe in the AL with the broad R57C10-Gal4 driver. Although local neurons (LNs) likely contribute to AL labeling as well, we found evidence of projection neuron (PN) activation (fig. S19). The activation of these secondary neurons was clearer when we used the PN-specific GH146-Gal4 (figs. S18, S20, and S21). With GH146-Gal4, we also found photoconverted PN axons projecting to the mushroom body (MB) calyx (CA) and the lateral horn (LH) (figs. S20 and S21 and movie S6), confirming that CaMPARI enables trans-synaptic circuit mapping. The labeled axons in LH arborize primarily ventrally (fig. S21), consistent with previous reports (26). Thus, pan-neuronal expression enables large-scale unbiased ensemble mapping across the brain. Coregistration of active cells to standardized atlases or the use of specific drivers to identify activated neurons (27) could be a powerful strategy for mapping pathways involved in specific behaviors.
We reasoned that sensory-driven circuit mapping with CaMPARI could also be complemented by the use of optogenetic stimulation to activate circuits downstream of an arbitrary cellular point of entry. We expressed the red light–sensitive channelrhodopsin CsChrimson (28) in Ir84a receptor neurons (fig. S22) using a LexA driver while expressing CaMPARI pan-neuronally with a Gal4 driver. We then mapped active circuits throughout the brain during optogenetic stimulation of this single-receptor neuron type (Fig. 4G). In addition to the glomeruli in the antennal lobe, where axons of ORNs and dendrites of PNs synapse (Fig. 4H), we observed axons of PNs projecting to the posterior areas of the brain (Fig. 4, I to L). One axon commissure, the medial antennal lobe tract (mALT), targets both CA and LH (Fig. 4, I and K), whereas another commissure, the medial-lateral antennal lobe tract (mlALT), targets LH directly (Fig. 4, I to L). As with PAA odor-driven activation, photoconverted PN axons mostly target the ventral part of LH. In addition, we detected activity in the MB vertical lobe (VL, Fig. 4, J and L) and VLP (Fig. 4, I and K), which are putative postsynaptic partners of PNs (Fig. 4J and fig. S23). More specifically, the α′ and β′ MB lobes were more photoconverted than α and β (fig. S24, E, F, G, and I compared to D), consistent with previous reports (29, 30). Using the higher-affinity CaMPARI, we attempted to mark circuit components further downstream (Fig. 4, K and L, and fig. S25). Although we found similar patterns of active PNs and VLP neurons as before (Fig. 4K), we also found other putative downstream circuit components such as LH neurons (Fig. 4L, cyan arrows), mushroom body output neurons (Fig. 4, J and L, and fig. S24), and descending neuron projections to the ventral nerve cord (fig. S25). Activation of deeper layers of the circuit suggests that CsChrimson may be able to drive the firing rates (and thus Ca2+ levels) of sensory neurons and downstream neurons more potently than natural stimuli.
In summary, CaMPARI was validated in four preparations in three organisms. Recording neural circuit activity free from restraint creates opportunities for studying complex behaviors such as social interaction, courtship, and so forth. In the case of direction-selective L2/3 pyramidal cells in mouse V1, CaMPARI labeling is consistent with results obtained from imaging of GECIs (12) and small-molecule dyes (31), as well as electrical recordings (21). In freely moving Drosophila larvae, vibration-evoked signals were restricted to chordotonal, but not proprioceptive or nociceptive, sensory neurons; the results show a functional difference in sensitivity of two distinct subtypes of chordotonal axons to vibration. In adult Drosophila brains, PC during both naturalistic and optogenetic stimulation of olfactory sensory pathways labeled activity in the expected OSN/PN circuit and, further, labeled PN projections to the LH and MB, and potential quaternary descending neurons in the VNC. In freely swimming zebrafish, activation patterns are consistent with hindbrain premotor and spinal cord motor output, as well as sensory modality-specific responses, which will be followed up experimentally. In all in vivo test cases, CaMPARI validates known results and offers extensions not otherwise possible: free movement, whole-brain activity mapping, a permanent signal compatible with fixation, and follow-up experiments such as targeted electrical recordings, multichannel immunohistochemistry (allowing cell type identification), and isolation and genetic profiling of labeled cells. These properties facilitate “forward functional neuromics,” in which active neural populations are labeled during behavior, drug treatment, sensory stimulation, and so forth in an unbiased fashion, with subsequent validation and characterization of labeled cells.
CaMPARI represents a new class of genetically encoded indicator that enables light-gated integration of Ca2+ flux in vivo with a time resolution of seconds to hours, specified by the user. Converted signal is persistent, amenable to multitrial accumulation, and compatible with fixation for post hoc whole-brain (or targeted-area) activity reconstruction. CaMPARI signal is ratiometric, correcting for expression-level differences. The protein engineering principles applied here could be extended to allow permanent marking of cell states or analytes other than Ca2+ by fusion of circularly permuted Eos variants to various ligand-binding or sensor domains, similar to previous GFP-based fluorescence intensity sensors (3234).

Acknowledgments

This work was performed as part of the Janelia GENIE project. We thank L. Lavis, K. Svoboda, R. Kerr, J. Macklin, G. Jefferis, Y. Aso, S. Namiki, A. Nern, J. Strother, and members of the Jayaraman lab, Looger lab, Zlatic lab, Svoboda lab, and GENIE project for discussions; A. Hu, B. Shields, and H. White for assistance with cell culture; J. Osborne, M. Peek, and T. Tabachnik for help with hardware design and fabrication; Janelia Fly Facility, especially K. Hibbard and B. Sharp for fly husbandry; Janelia Imaging Facility, especially P. Hulamm for support; Janelia Molecular Biology core for DNA sequencing and virus production; G. Paez for DNA constructs; A. Wong for assistance with preliminary experiments in adult Drosophila; and T. Zhao for neuTube. The crystal structure of calcium-free CaMPARI has been deposited in the Protein Data Bank with the accession code 4OY4. DNA constructs, AAV particles, and transgenic Drosophila were deposited for distribution at Addgene (http://www.addgene.org), the University of Pennsylvania Vector Core (www.med.upenn.edu/gtp/vectorcore), and the Bloomington Drosophila Stock Center (http://flystocks.bio.indiana.edu), respectively. Please contact M. Ahrens ([email protected]) for access to transgenic zebrafish.

Supplementary Material

Summary

Materials and Methods
Supplementary Text
Figs. S1 to S35
Tables S1 to S3
References (3565)
Movies S1 to S8

Resources

File (1260922s1.mp4)
File (1260922s2.mp4)
File (1260922s3.mp4)
File (1260922s4.mp4)
File (1260922s5.mp4)
File (1260922s6.mp4)
File (1260922s7.mp4)
File (1260922s8.mp4)
File (fosque.sm.pdf)

References and Notes

1
Baker P. F., Hodgkin A. L., Ridgway E. B., Depolarization and calcium entry in squid giant axons. J. Physiol. 218, 709–755 (1971).
2
Tank D. W., Sugimori M., Connor J. A., Llinás R. R., Spatially resolved calcium dynamics of mammalian Purkinje cells in cerebellar slice. Science 242, 773–777 (1988).
3
Chen J. L., Andermann M. L., Keck T., Xu N. L., Ziv Y., Imaging neuronal populations in behaving rodents: Paradigms for studying neural circuits underlying behavior in the mammalian cortex. J. Neurosci. 33, 17631–17640 (2013).
4
Riemensperger T., Pech U., Dipt S., Fiala A., Optical calcium imaging in the nervous system of Drosophila melanogaster. Biochim. Biophys. Acta 1820, 1169–1178 (2012).
5
Tian L., Hires S. A., Looger L. L., Imaging neuronal activity with genetically encoded calcium indicators. Cold Spring Harbor Protoc. 2012, 647–656 (2012).
6
Dreosti E., Odermatt B., Dorostkar M. M., Lagnado L., A genetically encoded reporter of synaptic activity in vivo. Nat. Methods 6, 883–889 (2009).
7
Guzowski J. F., Timlin J. A., Roysam B., McNaughton B. L., Worley P. F., Barnes C. A., Mapping behaviorally relevant neural circuits with immediate-early gene expression. Curr. Opin. Neurobiol. 15, 599–606 (2005).
8
Fields R. D., Eshete F., Stevens B., Itoh K., Action potential-dependent regulation of gene expression: Temporal specificity in ca2+, cAMP-responsive element binding proteins, and mitogen-activated protein kinase signaling. J. Neurosci. 17, 7252–7266 (1997).
9
Sheng H. Z., Fields R. D., Nelson P. G., Specific regulation of immediate early genes by patterned neuronal activity. J. Neurosci. Res. 35, 459–467 (1993).
10
Tsien R. Y., Very long-term memories may be stored in the pattern of holes in the perineuronal net. Proc. Natl. Acad. Sci. U.S.A. 110, 12456–12461 (2013).
11
O’Connor D. H., Huber D., Svoboda K., Reverse engineering the mouse brain. Nature 461, 923–929 (2009).
12
Chen T. W., Wardill T. J., Sun Y., Pulver S. R., Renninger S. L., Baohan A., Schreiter E. R., Kerr R. A., Orger M. B., Jayaraman V., Looger L. L., Svoboda K., Kim D. S., Ultrasensitive fluorescent proteins for imaging neuronal activity. Nature 499, 295–300 (2013).
13
Akerboom J., Rivera J. D., Guilbe M. M., Malavé E. C., Hernandez H. H., Tian L., Hires S. A., Marvin J. S., Looger L. L., Schreiter E. R., Crystal structures of the GCaMP calcium sensor reveal the mechanism of fluorescence signal change and aid rational design. J. Biol. Chem. 284, 6455–6464 (2009).
14
Wiedenmann J., Ivanchenko S., Oswald F., Schmitt F., Röcker C., Salih A., Spindler K. D., Nienhaus G. U., EosFP, a fluorescent marker protein with UV-inducible green-to-red fluorescence conversion. Proc. Natl. Acad. Sci. U.S.A. 101, 15905–15910 (2004).
15
Hoi H., Matsuda T., Nagai T., Campbell R. E., Highlightable Ca2+ indicators for live cell imaging. J. Am. Chem. Soc. 135, 46–49 (2013).
16
Information on materials and methods is available on Science Online.
17
Freeman J., Vladimirov N., Kawashima T., Mu Y., Sofroniew N. J., Bennett D. V., Rosen J., Yang C. T., Looger L. L., Ahrens M. B., Mapping brain activity at scale with cluster computing. Nat. Methods 11, 941–950 (2014).
18
Vladimirov N., Mu Y., Kawashima T., Bennett D. V., Yang C. T., Looger L. L., Keller P. J., Freeman J., Ahrens M. B., Light-sheet functional imaging in fictively behaving zebrafish. Nat. Methods 11, 883–884 (2014).
19
Hummel T., Krukkert K., Roos J., Davis G., Klämbt C., Drosophila Futsch/22C10 is a MAP1B-like protein required for dendritic and axonal development. Neuron 26, 357–370 (2000).
20
Ohyama T., Jovanic T., Denisov G., Dang T. C., Hoffmann D., Kerr R. A., Zlatic M., High-throughput analysis of stimulus-evoked behaviors in Drosophila larva reveals multiple modality-specific escape strategies. PLOS ONE 8, e71706 (2013).
21
Niell C. M., Stryker M. P., Highly selective receptive fields in mouse visual cortex. J. Neurosci. 28, 7520–7536 (2008).
22
Ruta V., Datta S. R., Vasconcelos M. L., Freeland J., Looger L. L., Axel R., A dimorphic pheromone circuit in Drosophila from sensory input to descending output. Nature 468, 686–690 (2010).
23
Yao Z., Macara A. M., Lelito K. R., Minosyan T. Y., Shafer O. T., Analysis of functional neuronal connectivity in the Drosophila brain. J. Neurophysiol. 108, 684–696 (2012).
24
Ito K., Shinomiya K., Ito M., Armstrong J. D., Boyan G., Hartenstein V., Harzsch S., Heisenberg M., Homberg U., Jenett A., Keshishian H., Restifo L. L., Rössler W., Simpson J. H., Strausfeld N. J., Strauss R., Vosshall L. B.Insect Brain Name Working Group, A systematic nomenclature for the insect brain. Neuron 81, 755–765 (2014).
25
Stensmyr M. C., Dweck H. K., Farhan A., Ibba I., Strutz A., Mukunda L., Linz J., Grabe V., Steck K., Lavista-Llanos S., Wicher D., Sachse S., Knaden M., Becher P. G., Seki Y., Hansson B. S., A conserved dedicated olfactory circuit for detecting harmful microbes in Drosophila. Cell 151, 1345–1357 (2012).
26
Grosjean Y., Rytz R., Farine J. P., Abuin L., Cortot J., Jefferis G. S., Benton R., An olfactory receptor for food-derived odours promotes male courtship in Drosophila. Nature 478, 236–240 (2011).
27
Jenett A., Rubin G. M., Ngo T. T., Shepherd D., Murphy C., Dionne H., Pfeiffer B. D., Cavallaro A., Hall D., Jeter J., Iyer N., Fetter D., Hausenfluck J. H., Peng H., Trautman E. T., Svirskas R. R., Myers E. W., Iwinski Z. R., Aso Y., DePasquale G. M., Enos A., Hulamm P., Lam S. C., Li H. H., Laverty T. R., Long F., Qu L., Murphy S. D., Rokicki K., Safford T., Shaw K., Simpson J. H., Sowell A., Tae S., Yu Y., Zugates C. T., A GAL4-driver line resource for Drosophila neurobiology. Cell Reports 2, 991–1001 (2012).
28
Klapoetke N. C., Murata Y., Kim S. S., Pulver S. R., Birdsey-Benson A., Cho Y. K., Morimoto T. K., Chuong A. S., Carpenter E. J., Tian Z., Wang J., Xie Y., Yan Z., Zhang Y., Chow B. Y., Surek B., Melkonian M., Jayaraman V., Constantine-Paton M., Wong G. K., Boyden E. S., Independent optical excitation of distinct neural populations. Nat. Methods 11, 338–346 (2014).
29
Turner G. C., Bazhenov M., Laurent G., Olfactory representations by Drosophila mushroom body neurons. J. Neurophysiol. 99, 734–746 (2008).
30
Lin A. C., Bygrave A. M., de Calignon A., Lee T., Miesenböck G., Sparse, decorrelated odor coding in the mushroom body enhances learned odor discrimination. Nat. Neurosci. 17, 559–568 (2014).
31
Ohki K., Chung S., Ch’ng Y. H., Kara P., Reid R. C., Functional imaging with cellular resolution reveals precise micro-architecture in visual cortex. Nature 433, 597–603 (2005).
32
Marvin J. S., Borghuis B. G., Tian L., Cichon J., Harnett M. T., Akerboom J., Gordus A., Renninger S. L., Chen T. W., Bargmann C. I., Orger M. B., Schreiter E. R., Demb J. B., Gan W. B., Hires S. A., Looger L. L., An optimized fluorescent probe for visualizing glutamate neurotransmission. Nat. Methods 10, 162–170 (2013).
33
St-Pierre F., Marshall J. D., Yang Y., Gong Y., Schnitzer M. J., Lin M. Z., High-fidelity optical reporting of neuronal electrical activity with an ultrafast fluorescent voltage sensor. Nat. Neurosci. 17, 884–889 (2014).
34
Tantama M., Martínez-François J. R., Mongeon R., Yellen G., Imaging energy status in live cells with a fluorescent biosensor of the intracellular ATP-to-ADP ratio. Nat. Commun. 4, 2550 (2013).
35
McKinney S. A., Murphy C. S., Hazelwood K. L., Davidson M. W., Looger L. L., A bright and photostable photoconvertible fluorescent protein. Nat. Methods 6, 131–133 (2009).
36
Paez-Segala M. G., et al., Nat. Methods (2015).
37
Studier F. W., Protein production by auto-induction in high density shaking cultures. Protein Expr. Purif. 41, 207–234 (2005).
38
Akerboom J., Carreras Calderón N., Tian L., Wabnig S., Prigge M., Tolö J., Gordus A., Orger M. B., Severi K. E., Macklin J. J., Patel R., Pulver S. R., Wardill T. J., Fischer E., Schüler C., Chen T. W., Sarkisyan K. S., Marvin J. S., Bargmann C. I., Kim D. S., Kügler S., Lagnado L., Hegemann P., Gottschalk A., Schreiter E. R., Looger L. L., Genetically encoded calcium indicators for multi-color neural activity imaging and combination with optogenetics. Front. Mol. Neurosci. 6, 2 (2013).
39
Kremers G. J., Piston D., Photoconversion of purified fluorescent proteins and dual-probe optical highlighting in live cells. J. Vis. Exp. 40, 1995 (2010).
40
Mütze J., Iyer V., Macklin J. J., Colonell J., Karsh B., Petrášek Z., Schwille P., Looger L. L., Lavis L. D., Harris T. D., Excitation spectra and brightness optimization of two-photon excited probes. Biophys. J. 102, 934–944 (2012).
41
Battye T. G., Kontogiannis L., Johnson O., Powell H. R., Leslie A. G., iMOSFLM: A new graphical interface for diffraction-image processing with MOSFLM. Acta Crystallogr. D Biol. Crystallogr. 67, 271–281 (2011).
42
A. G. W. Leslie, in Joint CCP4 + ESF–EAMCB Newsletter on Protein Crystallography (1992), vol. 26.
43
Winn M. D., Ballard C. C., Cowtan K. D., Dodson E. J., Emsley P., Evans P. R., Keegan R. M., Krissinel E. B., Leslie A. G., McCoy A., McNicholas S. J., Murshudov G. N., Pannu N. S., Potterton E. A., Powell H. R., Read R. J., Vagin A., Wilson K. S., Overview of the CCP4 suite and current developments. Acta Crystallogr. D Biol. Crystallogr. 67, 235–242 (2011).
44
McCoy A. J., Grosse-Kunstleve R. W., Adams P. D., Winn M. D., Storoni L. C., Read R. J., Phaser crystallographic software. J. Appl. Cryst. 40, 658–674 (2007).
45
Nienhaus K., Nienhaus G. U., Wiedenmann J., Nar H., Structural basis for photo-induced protein cleavage and green-to-red conversion of fluorescent protein EosFP. Proc. Natl. Acad. Sci. U.S.A. 102, 9156–9159 (2005).
46
Emsley P., Lohkamp B., Scott W. G., Cowtan K., Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).
47
Wardill T. J., Chen T. W., Schreiter E. R., Hasseman J. P., Tsegaye G., Fosque B. F., Behnam R., Shields B. C., Ramirez M., Kimmel B. E., Kerr R. A., Jayaraman V., Looger L. L., Svoboda K., Kim D. S., A neuron-based screening platform for optimizing genetically-encoded calcium indicators. PLOS ONE 8, e77728 (2013).
48
Akerboom J., Chen T. W., Wardill T. J., Tian L., Marvin J. S., Mutlu S., Calderón N. C., Esposti F., Borghuis B. G., Sun X. R., Gordus A., Orger M. B., Portugues R., Engert F., Macklin J. J., Filosa A., Aggarwal A., Kerr R. A., Takagi R., Kracun S., Shigetomi E., Khakh B. S., Baier H., Lagnado L., Wang S. S., Bargmann C. I., Kimmel B. E., Jayaraman V., Svoboda K., Kim D. S., Schreiter E. R., Looger L. L., Optimization of a GCaMP calcium indicator for neural activity imaging. J. Neurosci. 32, 13819–13840 (2012).
49
Pologruto T. A., Sabatini B. L., Svoboda K., ScanImage: Flexible software for operating laser scanning microscopes. Biomed. Eng. Online 2, 13 (2003).
50
Brainard D. H., The Psychophysics Toolbox. Spat. Vis. 10, 433–436 (1997).
51
Pelli D. G., The VideoToolbox software for visual psychophysics: Transforming numbers into movies. Spat. Vis. 10, 437–442 (1997).
52
Kerlin A. M., Andermann M. L., Berezovskii V. K., Reid R. C., Broadly tuned response properties of diverse inhibitory neuron subtypes in mouse visual cortex. Neuron 67, 858–871 (2010).
53
M. Westerfield, Eugene, OR, 1993.
54
Lister J. A., Robertson C. P., Lepage T., Johnson S. L., Raible D. W., nacre encodes a zebrafish microphthalmia-related protein that regulates neural-crest-derived pigment cell fate. Development 126, 3757–3767 (1999).
55
White R. M., Sessa A., Burke C., Bowman T., LeBlanc J., Ceol C., Bourque C., Dovey M., Goessling W., Burns C. E., Zon L. I., Transparent adult zebrafish as a tool for in vivo transplantation analysis. Cell Stem Cell 2, 183–189 (2008).
56
Urasaki A., Morvan G., Kawakami K., Functional dissection of the Tol2 transposable element identified the minimal cis-sequence and a highly repetitive sequence in the subterminal region essential for transposition. Genetics 174, 639–649 (2006).
57
Sato T., Takahoko M., Okamoto H., HuC:Kaede, a useful tool to label neural morphologies in networks in vivo. Genesis 44, 136–142 (2006).
58
Fisher S., Grice E. A., Vinton R. M., Bessling S. L., Urasaki A., Kawakami K., McCallion A. S., Evaluating the biological relevance of putative enhancers using Tol2 transposon-mediated transgenesis in zebrafish. Nat. Protoc. 1, 1297–1305 (2006).
59
Seelig J. D., Chiappe M. E., Lott G. K., Dutta A., Osborne J. E., Reiser M. B., Jayaraman V., Two-photon calcium imaging from head-fixed Drosophila during optomotor walking behavior. Nat. Methods 7, 535–540 (2010).
60
von Reyn C. R., Breads P., Peek M. Y., Zheng G. Z., Williamson W. R., Yee A. L., Leonardo A., Card G. M., A spike-timing mechanism for action selection. Nat. Neurosci. 17, 962–970 (2014).
61
Smear M., Shusterman R., O’Connor R., Bozza T., Rinberg D., Perception of sniff phase in mouse olfaction. Nature 479, 397–400 (2011).
62
Schindelin J., Arganda-Carreras I., Frise E., Kaynig V., Longair M., Pietzsch T., Preibisch S., Rueden C., Saalfeld S., Schmid B., Tinevez J. Y., White D. J., Hartenstein V., Eliceiri K., Tomancak P., Cardona A., Fiji: An open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).
63
Peng H., Ruan Z., Long F., Simpson J. H., Myers E. W., V3D enables real-time 3D visualization and quantitative analysis of large-scale biological image data sets. Nat. Biotechnol. 28, 348–353 (2010).
64
Kim J., Zhao T., Petralia R. S., Yu Y., Peng H., Myers E., Magee J. C., mGRASP enables mapping mammalian synaptic connectivity with light microscopy. Nat. Methods 9, 96–102 (2012).
65
Adam V., Lelimousin M., Boehme S., Desfonds G., Nienhaus K., Field M. J., Wiedenmann J., McSweeney S., Nienhaus G. U., Bourgeois D., Structural characterization of IrisFP, an optical highlighter undergoing multiple photo-induced transformations. Proc. Natl. Acad. Sci. U.S.A. 105, 18343–18348 (2008).

(0)eLetters

eLetters is a forum for ongoing peer review. eLetters are not edited, proofread, or indexed, but they are screened. eLetters should provide substantive and scholarly commentary on the article. Neither embedded figures nor equations with special characters can be submitted, and we discourage the use of figures and equations within eLetters in general. If a figure or equation is essential, please include within the text of the eLetter a link to the figure, equation, or full text with special characters at a public repository with versioning, such as Zenodo. Please read our Terms of Service before submitting an eLetter.

Log In to Submit a Response

No eLetters have been published for this article yet.

Information & Authors

Information

Published In

Science
Volume 347 | Issue 6223
13 February 2015

Submission history

Received: 8 September 2014
Accepted: 16 January 2015
Published in print: 13 February 2015

Permissions

Request permissions for this article.

Acknowledgments

This work was performed as part of the Janelia GENIE project. We thank L. Lavis, K. Svoboda, R. Kerr, J. Macklin, G. Jefferis, Y. Aso, S. Namiki, A. Nern, J. Strother, and members of the Jayaraman lab, Looger lab, Zlatic lab, Svoboda lab, and GENIE project for discussions; A. Hu, B. Shields, and H. White for assistance with cell culture; J. Osborne, M. Peek, and T. Tabachnik for help with hardware design and fabrication; Janelia Fly Facility, especially K. Hibbard and B. Sharp for fly husbandry; Janelia Imaging Facility, especially P. Hulamm for support; Janelia Molecular Biology core for DNA sequencing and virus production; G. Paez for DNA constructs; A. Wong for assistance with preliminary experiments in adult Drosophila; and T. Zhao for neuTube. The crystal structure of calcium-free CaMPARI has been deposited in the Protein Data Bank with the accession code 4OY4. DNA constructs, AAV particles, and transgenic Drosophila were deposited for distribution at Addgene (http://www.addgene.org), the University of Pennsylvania Vector Core (www.med.upenn.edu/gtp/vectorcore), and the Bloomington Drosophila Stock Center (http://flystocks.bio.indiana.edu), respectively. Please contact M. Ahrens ([email protected]) for access to transgenic zebrafish.

Authors

Affiliations

Benjamin F. Fosque*
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Present address: Department of Biochemistry and Molecular Biology, University of Chicago, Chicago, IL 60637, USA.
Yi Sun*
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Hod Dana*
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Chao-Tsung Yang
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Tomoko Ohyama
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Michael R. Tadross
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Ronak Patel
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Marta Zlatic
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Douglas S. Kim
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Misha B. Ahrens
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Vivek Jayaraman
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Loren L. Looger
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.
Eric R. Schreiter [email protected]
Howard Hughes Medical Institute, Janelia Farm Research Campus, 19700 Helix Drive, Ashburn, VA 20147, USA.

Notes

*
These authors contributed equally to this work.
Corresponding author. E-mail: [email protected]

Metrics & Citations

Metrics

Article Usage

Altmetrics

Citations

Cite as

Export citation

Select the format you want to export the citation of this publication.

Cited by

  1. DNA Methylation in Long-Term Memory, Physiology, 40, 4, (374-386), (2025).http://doi.org/10.1152/physiol.00032.2024
    Crossref
  2. Paternal ethanol exposure alters offspring motor skills and behavior in a sex-dependent manner and modifies early growth response 1 expression in the medial prefrontal cortex, Journal of Affective Disorders, 381, (388-400), (2025).http://doi.org/10.1016/j.jad.2025.04.031
    Crossref
  3. Neural substrates of cold nociception in Drosophila larva, eLife, 12, (2025).http://doi.org/10.7554/eLife.91582.3
    Crossref
  4. Neural substrates of cold nociception in Drosophila larva, eLife, 12, (2025).http://doi.org/10.7554/eLife.91582
    Crossref
  5. Cutting-edge methodologies for tagging and tracing active neuronal coding in the brain, Current Opinion in Neurobiology, 92, (102997), (2025).http://doi.org/10.1016/j.conb.2025.102997
    Crossref
  6. Adenosine signaling in glia modulates metabolic state-dependent behavior in Drosophila, Cell Reports, 44, 6, (115765), (2025).http://doi.org/10.1016/j.celrep.2025.115765
    Crossref
  7. Corticothalamic neurons in motor cortex have a permissive role in motor execution, Nature Communications, 16, 1, (2025).http://doi.org/10.1038/s41467-025-59954-1
    Crossref
  8. Protein Engineering for Spatiotemporally Resolved Cellular Monitoring, Annual Review of Analytical Chemistry, 18, 1, (217-240), (2025).http://doi.org/10.1146/annurev-anchem-070124-035857
    Crossref
  9. Engram and behavior: How memory is stored in the brain, Neurobiology of Learning and Memory, 219, (108047), (2025).http://doi.org/10.1016/j.nlm.2025.108047
    Crossref
  10. Engrams across diseases: Different pathologies – unifying mechanisms?, Neurobiology of Learning and Memory, 219, (108036), (2025).http://doi.org/10.1016/j.nlm.2025.108036
    Crossref
  11. See more
Loading...

View Options

View options

PDF format

Download this article as a PDF file

Download PDF

Log in to view the full text

AAAS ID LOGIN

Loading institution options

AAAS login provides access to Science for AAAS Members, and access to other journals in the Science family to users who have purchased individual subscriptions.

More options

Register for free to read this article

As a service to the community, this article is available for free. Login or register for free to read this article.

Purchase this issue in print

Buy a single issue of Science for just $15 USD.

Figures

Tables

Multimedia

Share

Share

Copy the article link

Share on social media

ScienceAdviser

Get Science’s award-winning newsletter with the latest news, commentary, and research, free to your inbox daily.