Open Access
17 January 2024 Imaging different cell populations in the mouse olfactory bulb using the genetically encoded voltage indicator ArcLight
Lee Min Leong, Douglas A. Storace
Author Affiliations +
Abstract

Genetically encoded voltage indicators (GEVIs) are protein-based optical sensors that allow for measurements from genetically defined populations of neurons. Although in vivo imaging in the mammalian brain with early generation GEVIs was difficult due to poor membrane expression and low signal-to-noise ratio, newer and more sensitive GEVIs have begun to make them useful for answering fundamental questions in neuroscience. We discuss principles of imaging using GEVIs and genetically encoded calcium indicators, both useful tools for in vivo imaging of neuronal activity, and review some of the recent mechanistic advances that have led to GEVI improvements. We provide an overview of the mouse olfactory bulb (OB) and discuss recent studies using the GEVI ArcLight to study different cell types within the bulb using both widefield and two-photon microscopy. Specific emphasis is placed on using GEVIs to begin to study the principles of concentration coding in the OB, how to interpret the optical signals from population measurements in the in vivo brain, and future developments that will push the field forward.

Techniques that allow for monitoring the activity of genetically heterogeneous populations of neurons would facilitate our ability to understand the brain. First reported more than 70 years ago, optical measurements of brain activity offered a potentially powerful answer to this challenging goal.14 The persistent efforts of pioneering neuroscientists and the broader community have yielded remarkable improvements to the hardware and sensors required to perform imaging in the brain. Thanks to these advances, the ability to use points of light to signal the activity of nerve cells, imagined nearly a century ago, is now a commonly used technique in neuroscience.5 Here we discuss some of the techniques used to measure neural activity in neurons and review recent experiments using protein-based sensors of voltage to study the activity of different cell populations in the mouse olfactory bulb (OB).

1.

Organic Dyes Versus Genetically Encoded Sensors

Imaging experiments using voltage and calcium dyes have undergone rapid and extensive development and improvements since they were first described in 1968 (8-anilinonaphthalene-1-sulfonic acid) and 1975 (arsenazo III), respectively.1,6,7 Since then, the dye toolkit has undergone extensive development resulting in improvements leading to reduced phototoxicity, improved kinetics, and signal-to-noise ratio.712 For example, the voltage sensitive dyes di-4-ANEPPS and ANNINE-6plus have time constants on the order of 103 to 109  s, and linear relationships with voltage sensitivity that yield fractional changes >8%ΔF/F/100  mV.1318

However, two major challenges of using organic dyes are that they must be physically introduced into the preparation, and they label all cells approximately equally.19,20 The lack of cell-specificity challenged efforts to understand the functional contributions of different cell types, except in exceptional cases where a dye can be anatomically restricted.2129

A class of protein-based optical sensors have been developed as an alternative to organic dyes that allow for neuron-type-specific labeling, which can be used to determine the role of different neuron types within and across brain areas.3032 This review primarily focuses on sensor variants in which the dye has been replaced by an optical reporter (e.g., a fluorescent protein) fused to a sensing element (e.g., a voltage or calcium sensing domain).20,33,34 However, other approaches incorporate a hybrid approach where a genetically encoded protein interfaces with an externally applied organic dye.35,36

The gene encoding the combined protein can be introduced into specific cell types using different promoters, viral vectors, and transgenic mice. Many different kinds of genetically encoded indicators have recently become available that report different cellular signals and binding of different neurotransmitters.3748 Our review discusses sensors of voltage and calcium, with a specific focus on the use of genetically encoded voltage indicators (GEVIs) to study the mammalian OB.

2.

Genetically Encoded Calcium and Voltage Indicators

Genetically encoded calcium indicators (GECIs) are protein-based sensors that report changes in intracellular calcium.33 Calcium imaging is often used as a proxy for changes in spiking activity because action potentials cause an increase in intracellular calcium. However, intracellular calcium dynamics are shaped by other sources, including subthreshold depolarization,49,50 through ligand-gated receptors,51,52 intracellular sequestration,5355 and dendritic coincidence detection.16 Intracellular calcium dynamics are also substantially slower than voltage changes and cannot perfectly recapitulate action potential activity except at lower rates of spiking activity.5658 The biophysical properties of genetically encoded indicators can further complicate the interpretation of an optical measurement. The speed in which a GECI transforms changes in intracellular calcium into fluorescence changes can contribute additional temporal filtering,5962 and in some cases, the kinetics have calcium concentration-dependent nonlinearities.63,64 Many GECIs also exhibit nonlinearities in their relationship between calcium binding and fluorescence changes with Hill coefficients (which describes the degree of cooperativity in ligand–receptor binding) as great as 3.59,60,62,64,65 Each GECI has a different calcium affinity, which defines the center of the calcium range they can detect.66 For sensors with linear relationships, this affinity has a straightforward impact on the interpretation of the calcium signals; the same absolute change in calcium will evoke the same magnitude fluorescence change. The interpretation of a nonlinear GECI measurement is more complex because the same change in intracellular calcium could result in different fluorescence changes depending on the calcium concentration. In principle, measurements could be restricted to the linear concentration range of a GECI, although this is difficult or impractical to measure in most preparations. However, ground truth experiments that combined intracellular electrode recordings and calcium imaging have shown that some GECIs have approximately linear relationships between action potential activity and calcium signals across some spike rates.65,67 Since their introduction 25 years ago, engineering efforts have improved GECI sensitivity, brightness, and speed.32,65,66,68,69 Importantly, the growing toolkit of widely available viral vectors and transgenic reporter mice have simplified their use.70,71 These developments have revolutionized neuroscience by facilitating the ability to measure changes in neural signaling in a variety of preparations.

If the experimental goal is to measure subthreshold events or changes in spiking activity, direct optical measurements of voltage changes would be ideal. GEVIs are protein-based sensors that optically report changes in membrane potential.20,34 GEVI development has been slower in contrast with GECIs, which reflects the unique challenges involved in measuring cellular voltage changes. GEVIs must be sufficiently fast to detect voltage changes that can occur across millisecond timescales, and their expression must be restricted to the membrane because nonmembrane expression will reduce the signal-to-noise ratio by contributing fluorescence that is not voltage-dependent.19,72 Several biophysical properties of a GEVI are important in evaluating its ability to optically report voltage signals in neural tissue and in the interpretation of its optical signals.7375 The voltage-sensitivity range of a GEVI determines the absolute voltage changes that will be translated into optical changes.7577 The sensitivity of a GEVI is defined as the dynamic range of the fluorescence change across its voltage-sensitivity range, which is a key determinant of the signal-to-noise ratio. GEVI brightness is important because relatively few protein molecules can be introduced into the membrane and rapid voltage changes require fast frame rates that limit photon integration time. A sensitive GEVI with very low baseline fluorescence will be difficult to measure due to signal-to-noise issues and may require intense or biologically incompatible illumination levels.78 The speed at which voltage changes are transformed into fluorescence changes defines the kinetics of a GEVI. Onset and offset kinetics are typically measured in response to a voltage step and reported as the time to reach its peak fluorescence change or decay back to its baseline. A highly sensitive GEVI with a slow onset will be unable to reach its peak fluorescence change in response to a fast voltage change.79 The fluorescence signal from a GEVI with a linear voltage-sensitivity slope centered around 20  mV with submillisecond kinetics will include similar proportions of subthreshold and action potential spiking activity. In comparison, the fluorescence from a GEVI with similar voltage sensitivity but slower onset dynamics will likely reflect more subthreshold voltage signals.80 Shifting the voltage sensitivity of a GEVI to a more positive range will yield fluorescence signals that primarily report action potential activity.73,77 In principle, these properties could be optimized to engineer highly specific and targeted probes.74,76,81,82

3.

Principles of GEVI Mechanisms of Action

The first generation of GEVIs based on fusions between green fluorescent protein (GFP) and the voltage-sensing domain (VSD) of potassium or sodium channels did not function well in mammalian cells due to their poor membrane expression.31,72,8385 The discovery and integration of the voltage-sensing phosphatase of Ciona intestinalis (sea squirt) with various fluorescent proteins was a milestone in GEVI development.86,87 In this sensor design, optical changes are believed to occur via voltage-dependent movement of the S4 transmembrane domain, which causes movement of the fluorescent protein via its linking element with the VSD.88 The first design strategy using the Ciona intestinalis VSD utilized Förster resonance energy transfer (FRET) fluorescent protein pairs that are attached to the N and C terminus of the VSD, respectively; changes in membrane potential are reported as FRET changes.86,89,90 Another significant milestone in GEVI development was the serendipitous discovery that modifications to the fluorescent protein and its linking position with the VSD can modulate GEVI sensitivity. In a GEVI utilizing the Ciona intestinalis VSD with the eGFP variant super ecliptic pHluorin, a point mutation (A227D) resulted in a large increase in sensitivity.91 Further tuning of this mutation along the linking position between the VSD and fluorescent protein yielded the GEVI ArcLight, which exhibited a 25-fold increase in sensitivity in comparison with its unmutated version.91 ArcLight has a sigmoidal voltage sensitivity relationship with a v1/2 around 26  mV. In response to a 100 mV depolarizing step, ArcLight responds with a sensitivity of 40%ΔF/F, which can be best fit with a double exponential curve with fast onset and offset components of 10 and 50 ms, respectively.34,63,91 In practice, these dynamics are sufficiently fast to solve individual action potentials at upward of 60 Hz.73,91,92 Although ArcLight is based on a variant of eGFP, its brightness will be a complex function of many experimental factors including the preparation type and expression level.91,93

Newer GEVIs have incorporated the voltage-gated phosphatase gene discovered in different species with various combinatorial mutations with alternate fluorescent proteins including a circularly permutated GFP inserted into the S3 and S4 loops of the transmembrane segments.14,92,9499 Another developmental strategy has been to incorporate microbial rhodopsins, which make use of the voltage sensitivity and fluorescence of the retinal residing in the opsin protein at the membrane.78,100,101 A change in membrane potential results in a chemical change that alters the fluorescence of the opsin retinal complex.102 The quantum yield of the opsin retinal complex is low, resulting in weak fluorescence that requires very high illumination intensities.101,103 Some of these limitations have been overcome by fusing opsins with fluorescent proteins to form an FRET pair, resulting in substantially brighter fluorescence than the opsin alone.104107 One hybrid strategy incorporates rhodamine-based voltage dyes with a genetically encoded HaloTag to target the dye to specific neuronal populations. Another strategy based on enzymatic cleavage via photoinduced electron transfer of voltage dyes, where VoltageFluor remains dim until activated porcine liver esterase.108,109 These advances in GEVI engineering have facilitated their use in biological discoveries in different brain regions and preparation types.106,110116

4.

Organization and Circuitry of the Olfactory Bulb

The sense of smell is critical for many animals as it plays a key role in locating danger, food, and mates. In mice, olfactory receptor neurons (ORNs) express one out of a large number of olfactory receptor proteins (1000), each with a distinct affinity for an odor.117120 Different odor-concentration pairings evoke varying degrees of activity across the olfactory receptor population, resulting in a combinatorial code, which is transmitted into the OB.121,122 Each ORN type typically maps to one or two regions of OB neuropil called glomeruli.123126 Each glomerulus is innervated by the apical dendrites of a specific population of mitral and tufted cells (MTCs), which receive ORN input and project broadly to the rest of the brain.127,128 This input–output transformation is shaped by a complex synaptic network of anatomically and genetically heterogeneous interneurons.129,130 ORNs drive activity to MTCs through feed-forward and lateral pathways via interneurons that surround glomeruli that innervate their parent or neighboring glomeruli.129,131135 An additional lateral modulatory circuit is mediated via granule cells that make inhibitory connections on the lateral dendrites of MTCs.136,137 These networks are further shaped by presynaptic modulation of the ORN axon terminals134,138,139 and cortical feedback from other brain regions.140145

Defining the functional role(s) of the different populations of neurons within the OB remains an important step in understanding the function of the bulb in olfactory sensory processing.129,146 A recent strategy in addressing this question is to measure and compare the glomerular signals originating from different OB cell types.67,147149 Glomerular imaging is powerful because upward of 100 glomeruli can be simultaneously measured from the dorsal surface of the OB, each of which contains the processes of different cell types involved in the input–output transformation for a single olfactory receptor type.129,150,151 Glomerular measurements to date have reflected the average of different populations of neurons innervating a glomerulus, trading single-cell resolution for a population average. Neural activity measurements of population and single cells can be used to answer different kinds of questions; both have provided critical insight into the logic of olfactory sensory processing.21,148,149,152156

Glomerular signals have been measured from nonspecific neuronal populations by measuring changes in light reflectance, autofluorescence, and fluorescence from bath applied dyes.151,157165 Early efforts to perform cell-specific glomerular imaging were carried out by infusing calcium sensitive organic dyes into the nasal cavity of rodents, which results in the dye being transported to the ORN axon terminals in the OB where they can be imaged.21,23,26,166 ORNs have also been imaged using genetically encoded indicators of synaptic vesicle release and calcium.148,149,167169 Glomerular measurements have been carried out from MTCs and two interneuron populations using transgenic mice expressing a GECI under control of a cell-specific promoter, or that express Cre recombinase combined with Cre-dependent adeno associated virus (AAV) transduction or reporter transgenic mice.67,148,149,152,170174 Cell-specific glomerular measurements have been important in defining basic coding principles of how odor information is encoded across the OB input–output transformation,118,148,149,152,172,174,175 the plasticity of different cell types to perceptual experiences,168,176178 and testing models of olfactory sensory processing.152,170

Surprisingly, few of the studies examining sensory coding in the mouse OB have been carried out using voltage imaging.63,148,150,165,179 In addition to providing a direct readout of changes in membrane potential from genetically defined cell populations, voltage imaging can measure fast changes in neural activity that would be otherwise obscured when measured using slower imaging methods, and in principle, should be able to measure depolarizing and hyperpolarizing signals. Together, these advances may eventually provide clearer insight into fast temporal coding, and the nature of the inhibitory circuits within the OB. Here we review recent studies that used the GEVI ArcLight to perform glomerular imaging in the mouse OB.

5. ArcLight Measurements from Nonspecific Cell Populations in the Olfactory Bulb

The first GEVI recordings from the OB were conducted using ArcLight constitutively expressed using an adeno associated virus serotype 1 (AAV1).63 The AAV1 was designed to co-express a nuclear localized mCherry fluorescent protein to facilitate visualization of the labeled neurons. Injecting the AAV1 into the mouse OB resulted in membrane localized expression of ArcLight and intracellular expression of mCherry [Fig. 1(a), left]. ArcLight was expressed in a population of OB neurons that included MTCs, consistent with prior reports of the tropism of AAV1 [Fig. 1(a), right].180 In vivo measurements were made from the dorsal surface of the OB in anesthetized mice using widefield fluorescence microscopy. Odors were presented using a flow-dilution olfactometer and respiration was monitored using a piezosensor placed against the chest of the animal.181 Individual glomeruli were not evident in the in vivo baseline fluorescence likely because of the presence of diffuse fluorescence originating from ArcLight expressed in neuropil in deeper layers of the OB [Fig. 1(b), left]. Using a frame subtraction analysis in which the imaging frames prior to the odor were subtracted from those acquired during the odor revealed odor-evoked fluorescence changes in glomerular sized regions of interest across the dorsal bulb [Fig. 1(b), right panel]. The fluorescence signals from the glomerular sized regions of interest had a high signal-to-noise ratio that was similar from trial to trial and were often coupled to respiration [Fig. 1(c)]. The resting fluorescence was similarly stable across different trials within a recording session, showing an average decline of 1% across individual imaging trials that lasted for 8  s (0.6%±0.02, measurements of baseline fluorescence in 3 to 8 consecutive trials in six different preparations; illumination provided by a 150-W Opti Quip Xenon arc lamp). The fluorescence from the ArcLight expression was substantially brighter than the endogenous autofluorescence present in the uninjected hemibulb, allowing it to overshadow the intrinsic autofluorescence signal that can be measured using the same filter set [Fig. 1(d)].163,164 Furthermore, intrinsic optical signals measured as changes in light reflectance at 705 nm were not statistically different in the ArcLight injected and uninjected hemibulb [Fig. 1(e)]. Therefore, AAV1 transduction of ArcLight yielded sufficiently bright fluorescence for in vivo imaging that exhibited minimal photobleaching and phototoxicity.

Fig. 1

(a) ArcLight expression in the OB from histological images at high (left) and low (right) magnification. (b) ArcLight fluorescence in vivo from the OB (left) and a frame subtraction analysis illustrating peaks of activity following odor presentation (right). (c) Single-trial measurements aligned to the onset of respiration from the region of interest in panel (b). (d) Fluorescence measurements from ArcLight injected and uninjected hemibulbs. The inset illustrates the baseline fluorescence and regions of interest. (e) Intrinsic light reflectance measurements from injected and uninjected hemibulbs. The inset illustrates that the intrinsic glomerular signals were not significantly different in the ArcLight injected and uninjected hemibulbs. onl, olfactory nerve layer; gl, glomerular layer; epl, external plexiform layer; and mcl, mitral cell layer.

NPh_11_3_033402_f001.png

In a subset of the ArcLight injected preparations, an AAV1 expressing the GECI GCaMP3 or GCaMP6f was also injected into the opposite hemibulb of the same animal [Figs. 2(a) and 2(b), insets]. Single inspirations of odor resulted in ArcLight and GECI fluorescence changes that could be detected in single trials [Figs. 2(a) and 2(b)]. In trials where the mouse took multiple sniffs of the odor, the ArcLight signal returned most of the way to the baseline between each inspiration; respiratory coupled modulations were less prominent when measured with GECIs [Figs. 2(c) and 2(d), compare red, green, and blue traces]. ArcLight had sufficient temporal resolution to resolve temporal heterogeneity across the glomerular population.182 Caudal-lateral glomeruli tended to respond earlier following a single sniff, whereas rostral-medial glomeruli responded more slowly [Fig. 2(e)]. ArcLight signals were smaller than either GECI but had significantly faster onset kinetics, rise times, and decay times [Fig. 2(f)]. To compare the ability of ArcLight and both GECIs to track respiratory coupling, the respiratory frequency was measured using a piezosensor pressed against the mouse’s chest. The signals from ArcLight and both GECIs were normalized to control for differences in signal-to-noise ratio, and their power at the respiratory frequency was measured before and after odor presentation. Odor presentation resulted in a significant increase in power at the respiratory frequency for all three sensors, although ArcLight had a significantly larger change than either GECI [Fig. 2(g)]. The time course of the ArcLight signal is similar to the timing of action potential activity of mitral cells in anesthetized rodents, which spike during inspiration.183185 The higher temporal resolution of ArcLight may more faithfully represent the time course of respiratory coupled activity in the OB.

Fig. 2

Response to one (a, b) and two (c, d) inhalations of an odor from opposite hemibulbs in two preparations injected with (a, c) ArcLight and GCaMP3 and (b, d) ArcLight and GCaMP6f. The mean fluorescence and regions of interest for the traces for the two preparations are illustrated in the insets. (e) Frame subtraction from the ArcLight injected hemibulb at the timepoints illustrated by the gray bars in panel (a). (f) Temporal properties of the three sensors to one odor inhalation. (g) Power at the respiration frequency for the three sensors during multiple odor inhalations.

NPh_11_3_033402_f002.png

6.

ArcLight in Specific OB Cell Types

Glomerular measurements from cell-specific populations have been carried out with ArcLight expressed using Cre-dependent AAV transduction and using the Ai86 ArcLight reporter transgenic line, which requires an intersectional breeding strategy to drive co-expression of ArcLight, tTA, and Cre recombinase.70,71,148,179 ORN targeting was achieved by first mating Ai86 to the Camk2a-tTA transgenic line.70,186 The resulting offspring that expressed both ArcLight and tTA were subsequently mated to the OMP-Cre transgenic line.187 The offspring that expressed all three genes exhibited ArcLight fluorescence in the olfactory nerve layer and glomeruli [Fig. 3(a)].179 Selective expression of ArcLight in MTCs was achieved by injecting a Cre-dependent AAV into the OB of protocadherin21-Cre transgenic mice, which resulted in ArcLight fluorescence present in MTC bodies, and their lateral and apical dendrites [Fig. 3(b)].148,188 A Cre-dependent AAV was injected into the OB of TH-Cre transgenic mice to express ArcLight in a population of interneurons. ArcLight fluorescence was present in neurons that surrounded glomeruli along with their associated glomerular processes [Fig. 3(c)].135,150,170 ArcLight was also expressed in a population of OB interneurons driven by the intersection of the Camk2a and EMX promoters by mating the offspring from pairings between Ai86 and Camk2a-tTA to the EMX-Cre transgenic line. The offspring that expressed all three genes exhibited ArcLight fluorescence in a subset of periglomerular and granule cells189 [Fig. 3(d)].

Fig. 3

ArcLight expression in (a) ORNs, (b) MTCs, (c) TH interneurons, and (d) Camk2a-EMX interneurons. onl, olfactory nerve layer; gl, glomerular layer; epl, external plexiform layer; mcl, mitral cell layer; and gcl, granule cell layer.

NPh_11_3_033402_f003.png

7.

ArcLight Measurements From Olfactory Receptor Neuron Glomeruli

Olfactory perception must be determined by the pattern of olfactory receptors that are activated by an odor stimulus. Because each ORN glomerulus reflects the input from one receptor type, glomerular measurements provide a readout of how odors are encoded across the receptor population. Widefield fluorescence imaging was used to measure odor-evoked activity from ORNs in anesthetized mice. Individual glomeruli were more clearly visible in the mean fluorescence than in the constitutive AAV experiment, consistent with the expression of ArcLight being restricted to glomeruli [Fig. 4(a)]. Odors evoked fluorescence changes with a high signal-to-noise ratio that were often coupled to each inspiration of the stimulus [Fig. 4(b)]. A frame subtraction analysis revealed that different odors evoked distinct patterns of activity across the glomerular population [Fig. 4(c), columns]. The same odors presented at different concentrations also changed the glomerular patterns; increasing the concentration increased the amplitude of active glomeruli and recruited glomeruli that were not responsive at lower concentrations [Fig. 4(c), rows]. The combination of activated ORNs changes in response to different odors as well as different concentrations of the same odor; representations of odor identity and concentration are confounded at the input to the mouse OB.21 The effect of concentration on different ORNs is evident in the fluorescence versus time course signal measured from different glomeruli [Fig. 4(d)]. The steep concentration dependence is quantified by plotting the peak normalized amplitude versus concentration for each glomerulus [Fig. 4(e), rois from panel (c)] and a population of preparations [Fig. 4(f)]. The Hill coefficients of ORN measurements across many glomerulus-odor pairings in a population of preparations ranged between 1 and 2.2, which are consistent with previous ORN concentration response relationships in a variety of different preparations.21,118,190195 These results are consistent with a model in which ORNs are broadly tuned with a range of affinities for different odors.117 The steep concentration-response relationships of individual ORNs combined with their narrow dynamic range is the basis of changes in the ORN glomerular maps in response to an odor or concentration change. A threshold concentration of an odor will activate a subset of high affinity ORNs; increasing the concentration saturates high-affinity receptors and recruits inactive low affinity receptors. Importantly, the glomerular ORN voltage imaging results are consistent with similar measurements using other kinds of optical sensors.21,23,118,148,167

Fig. 4

Odor-evoked measurements in OMP-ArcLight transgenic mice. (a) Mean fluorescence in vivo. (b) Single-trial odor response from one glomerulus. (c) Frame subtraction maps in response to different odor-concentration pairings. The numbers at the top left of the bottom panels indicate the maximum ΔF/F. (d) Methyl valerate responses at different concentrations in two different glomeruli. (e, f) Normalized response amplitude versus odor concentration for the glomeruli in panel (c) and for a population of preparations.

NPh_11_3_033402_f004.png

8.

ArcLight Measurements From MTC Glomeruli

Humans and other animals exhibit concentration invariant olfactory perception, which refers to the ability to recognize an odor as the same across concentration changes.196200 The observation that the patterns of activated ORN glomeruli change in response to different odors, and concentration changes of the same odor has raised the question of where and how concentration invariance is generated in the brain.117 Experimental and modeling work supported the possibility that circuitry within the OB could be important for generating more stable olfactory representations.117,201,202 This possibility was tested using a two-color imaging strategy, in which the signals from the ORNs innervating OB glomeruli were compared with the signals from the MTCs innervating the same glomeruli. In these experiments, the calcium sensitive dye fura dextran was infused into the nasal cavity of Pcdh21-Cre transgenic mice that received a Cre-dependent ArcLight-expressing AAV injection. The resulting mice had their ORN glomeruli selectively labeled with fura dextran and MTCs with ArcLight.26,188 Because fura dextran has an absorption spectra that is distinct from ArcLight, the signals originating from the ORNs and the MTCs innervating the same glomeruli could be differentiated by changing the excitation wavelength.55,63,91,148

The ORN and MTC glomerular signals were imaged from anesthetized mice using widefield fluorescence microscopy in response to odors presented across 2 log units of concentration. Increasing the concentration of the same odor increased the number of activated ORN glomeruli and altered the combinatorial pattern [Fig. 5(a), input]. In comparison, the spatial pattern of activation measured from the MTC glomeruli changed less across the same concentration changes [Fig. 5(a), output]. The change in concentration dependence can be visualized by comparing the time versus fluorescence traces for ORN and MTC signals innervating different glomeruli [Fig. 5(b)]. The peak ORN and MTC responses were quantified across different concentrations to illustrate the transformation from a highly concentration-dependent ORN input to a MTC response with a less steep concentration-response relationship [Fig. 5(c)]. The different effects of odor concentration on ORN and MTC glomerular maps are qualitatively similar to previous glomerular ORN measurements and single-cell MTC measurements.21,148,167,179,203

Fig. 5

(a) Frame subtraction analysis from ORN (left) and MTC (right) glomeruli at five concentrations of the odor methyl valerate. (b) Fluorescence traces from the regions of interest illustrated in panel (a). (c) Normalized amplitude versus concentration for all the identified glomeruli from the preparation in panel (a).

NPh_11_3_033402_f005.png

The mechanism(s) underlying this transformation remain unclear. Presynaptic inhibition onto the ORNs could dynamically compress the ORN concentration dependence into a narrower range.134 However, presynaptic inhibition seems unlikely to be the dominant mechanism since the ORN maps were still concentration-dependent and comparisons of Hill coefficients measured from ORN somata and ORN glomeruli were not significantly different.118 A second proposed mechanism involves a global normalization across the glomerular population, a process that can rescale the overall output while maintaining activity across glomeruli at relatively similar levels.117,201 There exist multiple mechanisms that could support a normalization process within the OB. One candidate is a population of interneurons that have long-range lateral connections within the OB and are important in generating stable concentration response relationships in individual mitral cells.117,135,170,202

9.

ArcLight Measurements From TH-Expressing Glomeruli

Measurements from dopaminergic interneuron glomeruli support their role in the broad transmission of concentration-dependent ORN signals.170 We performed voltage imaging from the TH interneuron glomeruli to test this model of concentration sensitivity. A Cre-dependent ArcLight-expressing AAV was injected into the OB of TH-Cre transgenic mice [Fig. 3(c)]. Imaging was performed in anesthetized mice in response to different odor-concentration pairings using a widefield fluorescence microscope. A frame subtraction analysis demonstrated that different odors evoked distinct glomerular peaks of activity across the dorsal bulb [Fig. 6(a), columns]. Increasing the concentration of the same odors strongly influenced the amplitude and pattern of the glomerular activity maps [Fig. 6(a), rows]. The impact of concentration on individual glomeruli is illustrated by comparing the fluorescence time course measurements of four TH glomeruli at different concentrations [Fig. 6(b)]. The effect of concentration on TH expressing glomeruli was quantified by measuring their peak response at different concentrations. The TH-ArcLight concentration response relationships were overlaid with the measurements from ORNs and MTCs also carried out using ArcLight [Fig. 6(c), red, blue, and black].148,150,179 The signals from ORNs and TH glomeruli were similarly concentration-dependent, and the MTC glomeruli changed much less across the same concentration range. These results indicate that the concentration-dependence of TH glomeruli are more similar to ORNs, a result consistent with a previous calcium imaging study from dopamine-expressing glomeruli.170

Fig. 6

(a) Frame subtraction analysis of different odor-concentration pairings. The odor concentration and maximum ΔF/F for each panel is listed in the top left and bottom right of each panel, respectively. (b) Responses from four glomeruli in response to methyl valerate from panel (a). (c) Normalized response amplitude versus odor concentration for three TH-ArcLight preparations (blue) overlaid with concentration response data from ORN (red) and MTC glomeruli (black).

NPh_11_3_033402_f006.png

10.

Two-Photon Imaging with ArcLight in the OB

Brain tissue is highly scattering that poses interpretive challenges for experiments using GEVIs with widefield fluorescence imaging. The degree of scattering is influenced by multiple factors that include the optical properties of the tissue, the imaging optics, and wavelength. In experiments carried out using a photodiode array, light focused on a slice of brain tissue resulted in photons striking multiple detectors separated up to 100  μm.204 This results in blurring of signals that will reflect an average of the signals emitted from all the labeled processes.30,204,205 Therefore, scattering imposes a fundamental limitation on the spatial resolution that can be measured in densely labeled tissue. These issues are exacerbated when imaging from voltage sensors because the signals originate from all parts of a cell. Sparser labeling can facilitate the interpretation of widefield fluorescence measurements by trading off the ability to record from many cells. The value of sparse labeling in voltage imaging is highlighted by experiments in which a voltage sensitive dye is restricted to a single cell, allowing high spatial and temporal resolution measurements from different processes within the same neuron.30,206208 Sparser expression has been achieved with GEVIs using low titer AAV transduction and incorporating targeting sequences to restrict expression to parts of a cell.70,81,98,110,209,210 Another strategy has been to use two-photon microscopy, a laser scanning technique that reduces light scattering by localizing the fluorescence emission to a small spatial volume, reducing the fluorescence that originates from out of focus areas, which substantially improves the lateral and axial resolution that can be achieved in scattering tissue.211,212 Early efforts illustrated the feasibility of incorporating GEVIs with two-photon imaging, although the measurements were limited by signal-to-noise considerations.213,214 Because not all GEVIs work similarly well under widefield fluorescence and two-photon microscopy, it is important to validate each GEVI using both imaging techniques.96,215

We tested the ability of ArcLight in reporting odor-evoked activity in the in vivo mouse OB using two-photon microscopy. Imaging was carried out following an injection of a constitutive AAV1 that expressed ArcLight in the OB. In comparison with widefield imaging experiments that used the same AAV1, individual glomeruli and neighboring processes could be clearly resolved in the mean fluorescence using two-photon imaging [Fig. 7(a), red polygon]. In experiments using slow frame rates (3  Hz), odor responses could be detected across a range of concentrations, although no respiratory coupling was observed [Fig. 7(b)]. Similar measurements done at a higher sampling rate (100 Hz) from a smaller imaging area [Fig. 7(a), red rectangle] were able to capture respiratory coupled oscillations [Fig. 7(c), taken from red rectangle in panel (a)].

Fig. 7

(a) In vivo two-photon mean fluorescence following constitutive AAV1 transduction of ArcLight. Odor responses from the regions of interest in panel (a) measured at (b) 3  Hz and (c) 100 Hz. Baseline fluorescence from the EMX-ArcLight transgenic mouse measured with (d) widefield fluorescence and (e) two-photon microscopy. The right panel was imaged at a higher magnification zoom from the box in the middle panel. (f) Odor evoked signals from the regions of interest in panels (d) and (e).

NPh_11_3_033402_f007.png

In EMX-ArcLight transgenic mice, widefield fluorescence signals were diffuse, consistent with the presence of ArcLight in deeper layers that contribute out of focus fluorescence [Fig. 7(d), histology in Fig. 3(d)]. In comparison, individual glomeruli and neighboring processes could be clearly visualized using two-photon microscopy [Fig. 7(d) versus Fig. 7(e)]. Odor responses were detected using widefield fluorescence and two-photon imaging [Fig. 7(f), left versus middle panel]. Respiratory coupled oscillations were detected in the widefield fluorescence data that were sampled at 125 Hz and in the two-photon imaging data that were performed at 31  Hz [Fig. 7(f), compare left and middle panels]. Using two-photon imaging, odor responses could be detected in small regions of interest that were consistent with the morphology of individual neurons [Fig. 7(e)]. The odor responses were not present everywhere in the imaging field of view, highlighting the higher spatial resolution afforded by the use of two-photon microscopy [Fig. 7(f), compare ROI1 and ROI2 in the middle and right panels]. Therefore, ArcLight reports odor-evoked activity in the mouse OB using both widefield fluorescence and two-photon imaging. These results also highlight that slower two-photon imaging scanning speeds can limit the ability to capture fast events [compare Fig. 7(b) to Figs. 7(c) and 7(f)]. Resonant-based scanners have increased the full frame imaging speed over earlier systems, although additional increases in speed come at a cost of reduced spatial resolution and signal-to-noise since faster scanning shortens the laser dwell time, which shortens the time during which emitted photons can be measured. Recent advances in the use of acousto-optic modulators that allow for arbitrary imaging patterns at high temporal resolution offer a promising alternative to current scanners to capture higher frequency neuronal signaling.98,110,216

11.

Interpretation of the Population Measurements

The interpretation of the optical signals measured using ArcLight is complicated and depends on the targeting strategy and imaging modality. Although ArcLight was targeted to specific OB cell populations in these experiments, the expression was present in many neurons. Consider OMP-ArcLight, which is less impacted by the complications of light scattering because the fluorescence was restricted to ORN glomeruli on the dorsal surface [Fig. 3(a)]. Ignoring the impact of scattered fluorescence from lateral glomeruli, the signal from each glomerulus reflects the input from many individual ORNs and therefore the measurement from a glomerulus is an average of all the active and inactive labeled neurons. If 1000 asynchronously active neurons in a volume of tissue express a GEVI, the ΔF from any single-active neuron will be overshadowed by the fluorescence of the inactive cells and reduced proportionally by the number of inactive neurons. Moreover, any hyperpolarized neurons would fluoresce in the opposite direction of the depolarized neurons, which would further diminish the averaged fluorescence change. Therefore, the signal from a single-action potential would depend on the number of synchronously active and inactive cells, and therefore measurements of action potential activity in population measurements likely requires precise temporal synchronization of spiking activity across many of the labeled neurons. Indeed, OB spiking activity is highly coupled to respiration in anesthetized mice, which provides a mechanism to tightly synchronize action potentials.154,183 However, the maximum ΔF/F response of ArcLight is only measured with a constant depolarization of 10 ms or longer.63,91 Since the voltage-dependence range of ArcLight includes subthreshold potentials, the population signals measured from ArcLight likely reflect some complex mixture of subthreshold and action potential activity. The use of GEVIs tuned to different voltage-ranges would allow for a clearer interpretation as to whether a population signal primarily reflects hyperpolarizing, subthreshold, or voltage changes.

Recent efforts to shape the voltage-response relationships of some GEVIs illustrate that this property can be modified with appropriate mutations.74,82,92 Ultimately there may not be a perfect GEVI for all applications, but a series of differently tuned GEVIs optimized to measure different kinds of voltage signals (e.g., subthreshold versus action potential events), analogous to the current strategies used for GECI development.65,66,69,82,217 For example, a GEVI with high sensitivity and instantaneous onset kinetics but with a voltage-dependence tuned to action potentials and a slow decay rate would be ideal for two-photon imaging on commonly available systems. A slower decay rate would extend the integration time to measure an action potential at a cost of losing the ability to resolve individual action potentials from exceptionally fast spiking neurons. The parallel development of sensitive GEVIs available in a range of fluorophores will facilitate the ability to perform simultaneous imaging from multiple sensors in a single imaging plane allowing for measurements of subthreshold and action potential activity, from different cell types or different neuronal signaling molecules.82,107

12.

Conclusions

Here we reviewed recent work demonstrating that the GEVI ArcLight can be used to measure glomerular signals from odor-evoked activity in the mouse OB from different cell types using widefield fluorescence and two-photon imaging. The faster temporal dynamics of ArcLight more precisely captured the respiratory coupled dynamics of glomerular activity when compared to GCaMP3 and GCaMP6f (Fig. 2). The ArcLight measurements illustrated that concentration coding is transformed across the OB circuit. These voltage imaging experiments validate work using GECIs demonstrating that a highly concentration-dependent ORN input signal is transformed into a more concentration invariant output response.148 Moreover, the glomerular output still maintained some sensitivity to concentration changes, indicating that concentration differences are still represented. In principle, such a coding scheme would allow downstream targets to recognize odors in a concentration invariant manner while still maintaining the ability to distinguish between concentration differences based on the output pattern of the bulb.

The growth of the number of laboratories developing and improving GEVIs over the last decade suggests an optimistic trajectory for making GEVIs easier to use in vivo, although the increasing number of those that are available complicates the choice of any particular one without clear comparisons of how they work in the in vivo brain. Moreover, the optimal GEVI choice likely depends on region-specific physiological parameters (e.g., varying spike widths and the overall balance of depolarization and hyperpolarization). Given that caveat, we propose that one potentially useful future goal is to have a model testing system for in vivo measurements by which all GEVIs could be screened for their ability to report population and single-cell activity using different microscopy techniques. We propose that the OB may be such a useful model brain area due to it containing dense neuropil within glomeruli as well as single cells that surround glomeruli that are easily accessible on the dorsal surface. Together, such developments and improvements will facilitate the uptake of GEVIs by the broader community to advance our understanding of the brain.

Disclosures

The authors declare no financial or scientific conflicts of interest.

Code and Data Availability

All data described in this manuscript are available from the authors upon request.

Acknowledgments

This work was supported by the National Institutes of Health (NIH), through a grant (no. DC020519) from the National Institute of Deafness and Communication Disorders.

References

1. 

I. Tasaki et al., “Changes in fluorescence, turbidity, and birefringence associated with nerve excitation,” Proc. Natl. Acad. Sci. U. S. A., 61 883 –888 https://doi.org/10.1073/pnas.61.3.883 (1968). Google Scholar

2. 

L. B. Cohen, R. D. Keynes and B. Hille, “Light scattering and birefringence changes during nerve activity,” Nature, 218 438 –441 https://doi.org/10.1038/218438a0 (1968). Google Scholar

3. 

D. K. Hill and R. D. Keynes, “Opacity changes in stimulated nerve,” J. Physiol., 108 278 –281 https://doi.org/10.1113/jphysiol.1949.sp004331 JPHYA7 0022-3751 (1949). Google Scholar

4. 

D. K. Hill, “The effect of stimulation on the opacity of a crustacean nerve trunk and its relation to fibre diameter,” J. Physiol., 111 283 –303 https://doi.org/10.1113/jphysiol.1950.sp004480 JPHYA7 0022-3751 (1950). Google Scholar

5. 

C. S. Sherrington, Man on His Nature, Cambridge University Press, Cambridge, UK (1937). Google Scholar

6. 

J. E. Brown, P. De Weer and B. M. Salzberg, “Optical measurement of changes in intracellular calcium,” Biophys. J., 118 788 –789 https://doi.org/10.1016/j.bpj.2019.06.034 BIOJAU 0006-3495 (2020). Google Scholar

7. 

J. E. Brown et al., “Rapid changes in intracellular free calcium concentration. Detection by metallochromic indicator dyes in squid giant axon,” Biophys. J., 15 1155 –1160 https://doi.org/10.1016/S0006-3495(75)85891-7 BIOJAU 0006-3495 (1975). Google Scholar

8. 

H. V. Davila et al., “A large change in axon fluorescence that provides a promising method for measuring membrane potential,” Nat. New Biol., 241 159 –160 https://doi.org/10.1038/newbio241159a0 (1973). Google Scholar

9. 

A. Grinvald, B. M. Salzberg and L. B. Cohen, “Simultaneous recording from several neurones in an invertebrate central nervous system,” Nature, 268 140 –142 https://doi.org/10.1038/268140a0 (1977). Google Scholar

10. 

B. M. Salzberg, H. V. Davila and L. B. Cohen, “Optical recording of impulses in individual neurones of an invertebrate central nervous system,” Nature, 246 508 –509 https://doi.org/10.1038/246508a0 (1973). Google Scholar

11. 

M. Canepari and D. Zecevic, Membrane Potential Imaging in the Nervous System: Methods and Applications, Springer( (2010). Google Scholar

12. 

O. Braubach, L. B. Cohen and Y. Choi, “Historical overview and general methods of membrane potential imaging,” Adv. Exp. Med. Biol., 859 3 –26 https://doi.org/10.1007/978-3-319-17641-3_1 AEMBAP 0065-2598 (2015). Google Scholar

13. 

M. Wachowiak et al., Brain Mapping: The Methods, 77 –95 Elsevier( (2002). Google Scholar

14. 

B. J. Baker et al., “Genetically encoded fluorescent voltage sensors using the voltage-sensing domain of Nematostella and Danio phosphatases exhibit fast kinetics,” J. Neurosci. Methods, 208 190 –196 https://doi.org/10.1016/j.jneumeth.2012.05.016 JNMEDT 0165-0270 (2012). Google Scholar

15. 

W. Frey et al., “Plasma membrane voltage changes during nanosecond pulsed electric field exposure,” Biophys. J., 90 3608 –3615 https://doi.org/10.1529/biophysj.105.072777 BIOJAU 0006-3495 (2006). Google Scholar

16. 

C. J. Roome and B. Kuhn, “Simultaneous dendritic voltage and calcium imaging and somatic recording from Purkinje neurons in awake mice,” Nat. Commun., 9 3388 https://doi.org/10.1038/s41467-018-05900-3 NCAOBW 2041-1723 (2018). Google Scholar

17. 

S. Pagès, D. Côté and P. De Koninck, “Optophysiological approach to resolve neuronal action potentials with high spatial and temporal resolution in cultured neurons,” Front. Cell. Neurosci., 5 20 https://doi.org/10.3389/fncel.2011.00020 (2011). Google Scholar

18. 

P. Fromherz et al., “ANNINE-6plus, a voltage-sensitive dye with good solubility, strong membrane binding and high sensitivity,” Eur. Biophys. J., 37 509 –514 https://doi.org/10.1007/s00249-007-0210-y EBJOE8 0175-7571 (2008). Google Scholar

19. 

B. J. Baker et al., “Genetically encoded fluorescent sensors of membrane potential,” Brain Cell Biol., 36 53 –67 https://doi.org/10.1007/s11068-008-9026-7 (2008). Google Scholar

20. 

D. Storace et al., “Genetically encoded protein sensors of membrane potential,” Adv. Exp. Med. Biol., 859 493 –509 https://doi.org/10.1007/978-3-319-17641-3_20 AEMBAP 0065-2598 (2015). Google Scholar

21. 

M. Wachowiak and L. B. Cohen, “Representation of odorants by receptor neuron input to the mouse olfactory bulb,” Neuron, 32 723 –735 https://doi.org/10.1016/S0896-6273(01)00506-2 NERNET 0896-6273 (2001). Google Scholar

22. 

J. V. Verhagen et al., “Sniffing controls an adaptive filter of sensory input to the olfactory bulb,” Nat. Neurosci., 10 631 –639 https://doi.org/10.1038/nn1892 NANEFN 1097-6256 (2007). Google Scholar

23. 

H. U. Fried, S. H. Fuss and S. I. Korsching, “Selective imaging of presynaptic activity in the mouse olfactory bulb shows concentration and structure dependence of odor responses in identified glomeruli,” Proc. Natl. Acad. Sci. U. S. A., 99 3222 –3227 https://doi.org/10.1073/pnas.052658399 (2002). Google Scholar

24. 

R. W. Friedrich and S. I. Korsching, “Combinatorial and chemotopic odorant coding in the zebrafish olfactory bulb visualized by optical imaging,” Neuron, 18 737 –752 https://doi.org/10.1016/S0896-6273(00)80314-1 NERNET 0896-6273 (1997). Google Scholar

25. 

R. W. Friedrich and S. I. Korsching, “Chemotopic, combinatorial, and noncombinatorial odorant representations in the olfactory bulb revealed using a voltage-sensitive axon tracer,” J. Neurosci., 18 9977 –9988 https://doi.org/10.1523/JNEUROSCI.18-23-09977.1998 JNRSDS 0270-6474 (1998). Google Scholar

26. 

S. Korsching, “Selective imaging of the receptor neuron population in the olfactory bulb of zebrafish and mice,” Chem. Sens., 30 (Suppl. 1), i101 –i102 https://doi.org/10.1093/chemse/bjh134 (2005). Google Scholar

27. 

J. B. Ackman, T. J. Burbridge and M. C. Crair, “Retinal waves coordinate patterned activity throughout the developing visual system,” Nature, 490 219 –225 https://doi.org/10.1038/nature11529 (2012). Google Scholar

28. 

Y. Tsau et al., “Dye screening and signal-to-noise ratio for retrogradely transported voltage-sensitive dyes,” J. Neurosci. Methods, 70 121 –129 https://doi.org/10.1016/S0165-0270(96)00109-4 JNMEDT 0165-0270 (1996). Google Scholar

29. 

P. Wenner et al., “Voltage-sensitive dye recording using retrogradely transported dye in the chicken spinal cord: staining and signal characteristics,” J. Neurosci. Methods, 70 111 –120 https://doi.org/10.1016/S0165-0270(96)00108-2 JNMEDT 0165-0270 (1996). Google Scholar

30. 

M. Zochowski et al., “Imaging membrane potential with voltage-sensitive dyes,” Biol. Bull., 198 1 –21 https://doi.org/10.2307/1542798 BIBUBX 0006-3185 (2000). Google Scholar

31. 

M. S. Siegel and E. Y. Isacoff, “A genetically encoded optical probe of membrane voltage,” Neuron, 19 735 –741 https://doi.org/10.1016/S0896-6273(00)80955-1 NERNET 0896-6273 (1997). Google Scholar

32. 

A. Miyawaki et al., “Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin,” Nature, 388 882 –887 https://doi.org/10.1038/42264 (1997). Google Scholar

33. 

S. A. Hires, L. Tian and L. L. Looger, “Reporting neural activity with genetically encoded calcium indicators,” Brain Cell Biol., 36 69 –86 https://doi.org/10.1007/s11068-008-9029-4 (2008). Google Scholar

34. 

D. Storace et al., “Toward better genetically encoded sensors of membrane potential,” Trends Neurosci., 39 277 –289 https://doi.org/10.1016/j.tins.2016.02.005 TNSCDR 0166-2236 (2016). Google Scholar

35. 

C. Deo et al., “The HaloTag as a general scaffold for far-red tunable chemigenetic indicators,” Nat. Chem. Biol., 17 718 –723 https://doi.org/10.1038/s41589-021-00775-w 1552-4450 (2021). Google Scholar

36. 

C. Deo and L. D. Lavis, “Synthetic and genetically encoded fluorescent neural activity indicators,” Curr. Opin. Neurobiol., 50 101 –108 https://doi.org/10.1016/j.conb.2018.01.003 COPUEN 0959-4388 (2018). Google Scholar

37. 

L. Duffet et al., “A genetically encoded sensor for in vivo imaging of orexin neuropeptides,” Nat. Methods, 19 231 –241 https://doi.org/10.1038/s41592-021-01390-2 1548-7091 (2022). Google Scholar

38. 

T. Patriarchi et al., “Ultrafast neuronal imaging of dopamine dynamics with designed genetically encoded sensors,” Science, 360 eaat4422 https://doi.org/10.1126/science.aat4422 SCIEAS 0036-8075 (2018). Google Scholar

39. 

H. Dong et al., “Genetically encoded sensors for measuring histamine release both in vitro and in vivo,” Neuron, 111 1564 –1576.e6 https://doi.org/10.1016/j.neuron.2023.02.024 NERNET 0896-6273 (2023). Google Scholar

40. 

T. Qian et al., “A genetically encoded sensor measures temporal oxytocin release from different neuronal compartments,” Nat. Biotechnol., 41 944 –957 https://doi.org/10.1038/s41587-022-01561-2 NABIF9 1087-0156 (2023). Google Scholar

41. 

A. Dong et al., “A fluorescent sensor for spatiotemporally resolved imaging of endocannabinoid dynamics in vivo,” Nat. Biotechnol., 40 787 –798 https://doi.org/10.1038/s41587-021-01074-4 NABIF9 1087-0156 (2022). Google Scholar

42. 

Z. Wu et al., “A sensitive GRAB sensor for detecting extracellular ATP in vitro and in vivo,” Neuron, 110 770 –782.e5 https://doi.org/10.1016/j.neuron.2021.11.027 NERNET 0896-6273 (2022). Google Scholar

43. 

J. Wan et al., “A genetically encoded sensor for measuring serotonin dynamics,” Nat. Neurosci., 24 746 –752 https://doi.org/10.1038/s41593-021-00823-7 NANEFN 1097-6256 (2021). Google Scholar

44. 

M. Jing et al., “An optimized acetylcholine sensor for monitoring in vivo cholinergic activity,” Nat. Methods, 17 1139 –1146 https://doi.org/10.1038/s41592-020-0953-2 1548-7091 (2020). Google Scholar

45. 

A. Aggarwal et al., “Glutamate indicators with improved activation kinetics and localization for imaging synaptic transmission,” Nat. Methods, 20 925 –934 https://doi.org/10.1038/s41592-023-01863-6 1548-7091 (2023). Google Scholar

46. 

J. S. Marvin et al., “Stability, affinity, and chromatic variants of the glutamate sensor iGluSnFR,” Nat. Methods, 15 936 –939 https://doi.org/10.1038/s41592-018-0171-3 1548-7091 (2018). Google Scholar

47. 

J. Wu et al., “Genetically encoded glutamate indicators with altered color and topology,” ACS Chem. Biol., 13 1832 –1837 https://doi.org/10.1021/acschembio.7b01085 (2018). Google Scholar

48. 

J. S. Marvin et al., “An optimized fluorescent probe for visualizing glutamate neurotransmission,” Nat. Methods, 10 162 –170 https://doi.org/10.1038/nmeth.2333 1548-7091 (2013). Google Scholar

49. 

S. Charpak et al., “Odor-evoked calcium signals in dendrites of rat mitral cells,” Proc. Natl. Acad. Sci. U. S. A., 98 1230 –1234 https://doi.org/10.1073/pnas.98.3.1230 (2001). Google Scholar

50. 

F. Helmchen et al., “In vivo dendritic calcium dynamics in deep-layer cortical pyramidal neurons,” Nat. Neurosci., 2 989 –996 https://doi.org/10.1038/14788 NANEFN 1097-6256 (1999). Google Scholar

51. 

J. Schiller et al., “NMDA spikes in basal dendrites of cortical pyramidal neurons,” Nature, 404 285 –289 https://doi.org/10.1038/35005094 (2000). Google Scholar

52. 

W. N. Ross, “Understanding calcium waves and sparks in central neurons,” Nat. Rev. Neurosci., 13 157 –168 https://doi.org/10.1038/nrn3168 NRNAAN 1471-003X (2012). Google Scholar

53. 

O. Garaschuk, Y. Yaari and A. Konnerth, “Release and sequestration of calcium by ryanodine-sensitive stores in rat hippocampal neurones,” J. Physiol., 502 (1), 13 –30 https://doi.org/10.1111/j.1469-7793.1997.013bl.x JPHYA7 0022-3751 (1997). Google Scholar

54. 

N. Emptage, T. V. Bliss and A. Fine, “Single synaptic events evoke NMDA receptor-mediated release of calcium from internal stores in hippocampal dendritic spines,” Neuron, 22 115 –124 https://doi.org/10.1016/S0896-6273(00)80683-2 NERNET 0896-6273 (1999). Google Scholar

55. 

L. D. Pozzo Miller et al., “Ca2+ release from intracellular stores induced by afferent stimulation of CA3 pyramidal neurons in hippocampal slices,” J. Neurophysiol., 76 554 –562 https://doi.org/10.1152/jn.1996.76.1.554 JONEA4 0022-3077 (1996). Google Scholar

56. 

J. C. Callaway and W. N. Ross, “Frequency-dependent propagation of sodium action potentials in dendrites of hippocampal CA1 pyramidal neurons,” J. Neurophysiol., 74 1395 –1403 https://doi.org/10.1152/jn.1995.74.4.1395 JONEA4 0022-3077 (1995). Google Scholar

57. 

H. Markram, P. J. Helm and B. Sakmann, “Dendritic calcium transients evoked by single back-propagating action potentials in rat neocortical pyramidal neurons,” J. Physiol., 485 (1), 1 –20 https://doi.org/10.1113/jphysiol.1995.sp020708 JPHYA7 0022-3751 (1995). Google Scholar

58. 

K. Svoboda et al., “In vivo dendritic calcium dynamics in neocortical pyramidal neurons,” Nature, 385 161 –165 https://doi.org/10.1038/385161a0 (1997). Google Scholar

59. 

X. R. Sun et al., “Fast GCaMPs for improved tracking of neuronal activity,” Nat. Commun., 4 2170 https://doi.org/10.1038/ncomms3170 NCAOBW 2041-1723 (2013). Google Scholar

60. 

A. Badura et al., “Fast calcium sensor proteins for monitoring neural activity,” Neurophotonics, 1 025008 https://doi.org/10.1117/1.NPh.1.2.025008 (2014). Google Scholar

61. 

T. Rose et al., “Putting a finishing touch on GECIs,” Front. Mol. Neurosci., 7 88 https://doi.org/10.3389/fnmol.2014.00088 (2014). Google Scholar

62. 

T. Thestrup et al., “Optimized ratiometric calcium sensors for functional in vivo imaging of neurons and T lymphocytes,” Nat. Methods, 11 175 –182 https://doi.org/10.1038/nmeth.2773 1548-7091 (2014). Google Scholar

63. 

D. A. Storace et al., “Monitoring brain activity with protein voltage and calcium sensors,” Sci. Rep., 5 10212 https://doi.org/10.1038/srep10212 SRCEC3 2045-2322 (2015). Google Scholar

64. 

J. Akerboom et al., “Optimization of a GCaMP calcium indicator for neural activity imaging,” J. Neurosci., 32 13819 –13840 https://doi.org/10.1523/JNEUROSCI.2601-12.2012 JNRSDS 0270-6474 (2012). Google Scholar

65. 

Y. Zhang et al., “Fast and sensitive GCaMP calcium indicators for imaging neural populations,” Nature, 615 884 –891 https://doi.org/10.1038/s41586-023-05828-9 (2023). Google Scholar

66. 

T. W. Chen et al., “Ultrasensitive fluorescent proteins for imaging neuronal activity,” Nature, 499 295 –300 https://doi.org/10.1038/nature12354 (2013). Google Scholar

67. 

M. Wachowiak et al., “Optical dissection of odor information processing in vivo using GCaMPs expressed in specified cell types of the olfactory bulb,” J. Neurosci., 33 5285 –5300 https://doi.org/10.1523/JNEUROSCI.4824-12.2013 JNRSDS 0270-6474 (2013). Google Scholar

68. 

J. Li et al., “Engineering of NEMO as calcium indicators with large dynamics and high sensitivity,” Nat. Methods, 20 918 –924 https://doi.org/10.1038/s41592-023-01852-9 1548-7091 (2023). Google Scholar

69. 

H. Dana et al., “High-performance calcium sensors for imaging activity in neuronal populations and microcompartments,” Nat. Methods, 16 649 –657 https://doi.org/10.1038/s41592-019-0435-6 1548-7091 (2019). Google Scholar

70. 

T. L. Daigle et al., “A suite of transgenic driver and reporter mouse lines with enhanced brain-cell-type targeting and functionality,” Cell, 174 465 –480.e22 https://doi.org/10.1016/j.cell.2018.06.035 CELLB5 0092-8674 (2018). Google Scholar

71. 

L. Madisen et al., “Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance,” Neuron, 85 942 –958 https://doi.org/10.1016/j.neuron.2015.02.022 NERNET 0896-6273 (2015). Google Scholar

72. 

B. J. Baker et al., “Three fluorescent protein voltage sensors exhibit low plasma membrane expression in mammalian cells,” J. Neurosci. Methods, 161 32 –38 https://doi.org/10.1016/j.jneumeth.2006.10.005 JNMEDT 0165-0270 (2007). Google Scholar

73. 

J. K. Rhee et al., “Biophysical parameters of GEVIs: considerations for imaging voltage,” Biophys. J., 119 1 –8 https://doi.org/10.1016/j.bpj.2020.05.019 BIOJAU 0006-3495 (2020). Google Scholar

74. 

A. Jung et al., “Linker length and fusion site composition improve the optical signal of genetically encoded fluorescent voltage sensors,” Neurophotonics, 2 021012 https://doi.org/10.1117/1.NPh.2.2.021012 (2015). Google Scholar

75. 

M. Sepehri Rad, L. B. Cohen and B. J. Baker, “Conserved amino acids residing outside the voltage field can shift the voltage sensitivity and increase the signal speed and size of Ciona based GEVIs,” Front. Cell Dev. Biol., 10 868143 https://doi.org/10.3389/fcell.2022.868143 (2022). Google Scholar

76. 

A. Jung et al., “Modulating the voltage-sensitivity of a genetically encoded voltage indicator,” Exp. Neurobiol., 26 241 –251 https://doi.org/10.5607/en.2017.26.5.241 (2017). Google Scholar

77. 

S. Lee et al., “Improving a genetically encoded voltage indicator by modifying the cytoplasmic charge composition,” Sci. Rep., 7 8286 https://doi.org/10.1038/s41598-017-08731-2 SRCEC3 2045-2322 (2017). Google Scholar

78. 

J. M. Kralj et al., “Electrical spiking in Escherichia coli probed with a fluorescent voltage-indicating protein,” Science, 333 345 –348 https://doi.org/10.1126/science.1204763 SCIEAS 0036-8075 (2011). Google Scholar

79. 

Y. Bando et al., “Comparative evaluation of genetically encoded voltage indicators,” Cell Rep., 26 802 –813.e4 https://doi.org/10.1016/j.celrep.2018.12.088 (2019). Google Scholar

80. 

M. Milosevic et al., “In vitro testing of voltage indicators: Archon1, ArcLightD, ASAP1, ASAP2s, ASAP3b, Bongwoori-Pos6, BeRST1, FlicR1 and Chi-VSFP-Butterfly,” eneuro, 7 ENEURO.0060-0020 https://doi.org/10.1523/ENEURO.0060-20.2020 (2020). Google Scholar

81. 

M. Sepehri Rad et al., “Monitoring voltage fluctuations of intracellular membranes,” Sci. Rep., 8 6911 https://doi.org/10.1038/s41598-018-25083-7 SRCEC3 2045-2322 (2018). Google Scholar

82. 

Y. Bando, M. Wenzel and R. Yuste, “Simultaneous two-photon imaging of action potentials and subthreshold inputs in vivo,” Nat. Commun., 12 7229 https://doi.org/10.1038/s41467-021-27444-9 NCAOBW 2041-1723 (2021). Google Scholar

83. 

G. Guerrero et al., “Tuning FlaSh: redesign of the dynamics, voltage range, and color of the genetically encoded optical sensor of membrane potential,” Biophys. J., 83 3607 –3618 https://doi.org/10.1016/S0006-3495(02)75361-7 BIOJAU 0006-3495 (2002). Google Scholar

84. 

K. Ataka and V. A. Pieribone, “A genetically targetable fluorescent probe of channel gating with rapid kinetics,” Biophys. J., 82 509 –516 https://doi.org/10.1016/S0006-3495(02)75415-5 BIOJAU 0006-3495 (2002). Google Scholar

85. 

R. Sakai et al., “Design and characterization of a DNA‐encoded, voltage‐sensitive fluorescent protein,” Eur. J. Neurosci., 13 2314 –2318 https://doi.org/10.1046/j.0953-816x.2001.01617.x EJONEI 0953-816X (2001). Google Scholar

86. 

D. Dimitrov et al., “Engineering and characterization of an enhanced fluorescent protein voltage sensor,” PLoS One, 2 e440 https://doi.org/10.1371/journal.pone.0000440 POLNCL 1932-6203 (2007). Google Scholar

87. 

Y. Murata et al., “Phosphoinositide phosphatase activity coupled to an intrinsic voltage sensor,” Nature, 435 1239 –1243 https://doi.org/10.1038/nature03650 (2005). Google Scholar

88. 

B. E. Kang and B. J. Baker, “Pado, a fluorescent protein with proton channel activity can optically monitor membrane potential, intracellular pH, and map gap junctions,” Sci. Rep., 6 23865 https://doi.org/10.1038/srep23865 SRCEC3 2045-2322 (2016). Google Scholar

89. 

U. Sung et al., “Developing fast fluorescent protein voltage sensors by optimizing FRET interactions,” PLoS One, 10 e0141585 https://doi.org/10.1371/journal.pone.0141585 POLNCL 1932-6203 (2015). Google Scholar

90. 

H. Tsutsui et al., “Improving membrane voltage measurements using FRET with new fluorescent proteins,” Nat. Methods, 5 683 –685 https://doi.org/10.1038/nmeth.1235 1548-7091 (2008). Google Scholar

91. 

L. Jin et al., “Single action potentials and subthreshold electrical events imaged in neurons with a fluorescent protein voltage probe,” Neuron, 75 779 –785 https://doi.org/10.1016/j.neuron.2012.06.040 NERNET 0896-6273 (2012). Google Scholar

92. 

H. H. Piao et al., “Combinatorial mutagenesis of the voltage-sensing domain enables the optical resolution of action potentials firing at 60 Hz by a genetically encoded fluorescent sensor of membrane potential,” J. Neurosci., 35 372 –385 https://doi.org/10.1523/JNEUROSCI.3008-14.2015 JNRSDS 0270-6474 (2015). Google Scholar

93. 

R. P. Ilagan et al., “A new bright green-emitting fluorescent protein: engineered monomeric and dimeric forms,” FEBS J., 277 1967 –1978 https://doi.org/10.1111/j.1742-4658.2010.07618.x 1742-464X (2010). Google Scholar

94. 

Z. Han et al., “Fluorescent protein voltage probes derived from ArcLight that respond to membrane voltage changes with fast kinetics,” PLoS One, 8 e81295 https://doi.org/10.1371/journal.pone.0081295 POLNCL 1932-6203 (2013). Google Scholar

95. 

Z. Han et al., “Mechanistic studies of the genetically encoded fluorescent protein voltage probe ArcLight,” PLoS One, 9 e113873 https://doi.org/10.1371/journal.pone.0113873 POLNCL 1932-6203 (2014). Google Scholar

96. 

S. Chamberland et al., “Fast two-photon imaging of subcellular voltage dynamics in neuronal tissue with genetically encoded indicators,” Elife, 6 e25690 https://doi.org/10.7554/eLife.25690 (2017). Google Scholar

97. 

F. St-Pierre et al., “High-fidelity optical reporting of neuronal electrical activity with an ultrafast fluorescent voltage sensor,” Nat. Neurosci., 17 884 –889 https://doi.org/10.1038/nn.3709 NANEFN 1097-6256 (2014). Google Scholar

98. 

Z. Liu et al., “Sustained deep-tissue voltage recording using a fast indicator evolved for two-photon microscopy,” Cell, 185 3408 –3425.e29 https://doi.org/10.1016/j.cell.2022.07.013 CELLB5 0092-8674 (2022). Google Scholar

99. 

J. Platisa et al., “High-speed low-light in vivo two-photon voltage imaging of large neuronal populations,” Nat. Methods, 20 1095 –1103 https://doi.org/10.1038/s41592-023-01820-3 1548-7091 (2023). Google Scholar

100. 

W. Junge and H. T. Witt, “On the ion transport system of photosynthesis—investigations on a molecular level,” Zeitschr. Naturforsch. B, 23 244 –254 https://doi.org/10.1515/znb-1968-0222 (1968). Google Scholar

101. 

J. M. Kralj et al., “Optical recording of action potentials in mammalian neurons using a microbial rhodopsin,” Nat. Methods, 9 90 –95 https://doi.org/10.1038/nmeth.1782 1548-7091 (2012). Google Scholar

102. 

D. Maclaurin et al., “Mechanism of voltage-sensitive fluorescence in a microbial rhodopsin,” Proc. Natl. Acad. Sci. U. S. A., 110 5939 –5944 https://doi.org/10.1073/pnas.1215595110 (2013). Google Scholar

103. 

D. R. Hochbaum et al., “All-optical electrophysiology in mammalian neurons using engineered microbial rhodopsins,” Nat. Methods, 11 825 –833 https://doi.org/10.1038/nmeth.3000 1548-7091 (2014). Google Scholar

104. 

Y. Gong et al., “Imaging neural spiking in brain tissue using FRET-opsin protein voltage sensors,” Nat. Commun., 5 3674 https://doi.org/10.1038/ncomms4674 NCAOBW 2041-1723 (2014). Google Scholar

105. 

P. Zou et al., “Bright and fast multicoloured voltage reporters via electrochromic FRET,” Nat. Commun., 5 4625 https://doi.org/10.1038/ncomms5625 NCAOBW 2041-1723 (2014). Google Scholar

106. 

M. Kannan et al., “Dual-polarity voltage imaging of the concurrent dynamics of multiple neuron types,” Science, 378 eabm8797 https://doi.org/10.1126/science.abm8797 SCIEAS 0036-8075 (2022). Google Scholar

107. 

M. Kannan et al., “Fast, in vivo voltage imaging using a red fluorescent indicator,” Nat. Methods, 15 1108 –1116 https://doi.org/10.1038/s41592-018-0188-7 1548-7091 (2018). Google Scholar

108. 

P. E. Deal et al., “Covalently tethered rhodamine voltage reporters for high speed functional imaging in brain tissue,” J. Am. Chem. Soc., 142 614 –622 https://doi.org/10.1021/jacs.9b12265 JACSAT 0002-7863 (2019). Google Scholar

109. 

P. Liu et al., “Fluorogenic targeting of voltage-sensitive dyes to neurons,” J. Am. Chem. Soc., 139 17334 –17340 https://doi.org/10.1021/jacs.7b07047 JACSAT 0002-7863 (2017). Google Scholar

110. 

V. Villette et al., “Ultrafast two-photon imaging of a high-gain voltage indicator in awake behaving mice,” Cell, 179 1590 –1608.e23 https://doi.org/10.1016/j.cell.2019.11.004 CELLB5 0092-8674 (2019). Google Scholar

111. 

G. Cao et al., “Genetically targeted optical electrophysiology in intact neural circuits,” Cell, 154 904 –913 https://doi.org/10.1016/j.cell.2013.07.027 CELLB5 0092-8674 (2013). Google Scholar

112. 

D. Raccuglia et al., “Presynaptic GABA receptors mediate temporal contrast enhancement in drosophila olfactory sensory neurons and modulate odor-driven behavioral kinetics,” eNeuro, 3 ENEURO.0080-16.2016 https://doi.org/10.1523/ENEURO.0080-16.2016 (2016). Google Scholar

113. 

D. Raccuglia et al., “Network-specific synchronization of electrical slow-wave oscillations regulates sleep drive in drosophila,” Curr. Biol., 29 3611 –3621.e3 https://doi.org/10.1016/j.cub.2019.08.070 CUBLE2 0960-9822 (2019). Google Scholar

114. 

A. J. Matzke and M. Matzke, “Expression and testing in plants of ArcLight, a genetically-encoded voltage indicator used in neuroscience research,” BMC Plant Biol., 15 245 https://doi.org/10.1186/s12870-015-0633-z (2015). Google Scholar

115. 

P. Y. Borden et al., “Genetically expressed voltage sensor ArcLight for imaging large scale cortical activity in the anesthetized and awake mouse,” Neurophotonics, 4 031212 https://doi.org/10.1117/1.NPh.4.3.031212 (2017). Google Scholar

116. 

J. Jimenez-Martin et al., “Cholinergic modulation of sensory processing in awake mouse cortex,” Sci. Rep., 11 17525 https://doi.org/10.1038/s41598-021-96696-8 SRCEC3 2045-2322 (2021). Google Scholar

117. 

T. A. Cleland et al., “Sequential mechanisms underlying concentration invariance in biological olfaction,” Front. Neuroeng., 4 21 https://doi.org/10.3389/fneng.2011.00021 (2011). Google Scholar

118. 

J. D. Zak et al., “Antagonistic odor interactions in olfactory sensory neurons are widespread in freely breathing mice,” Nat. Commun., 11 3350 https://doi.org/10.1038/s41467-020-17124-5 NCAOBW 2041-1723 (2020). Google Scholar

119. 

L. Buck and R. Axel, “A novel multigene family may encode odorant receptors: a molecular basis for odor recognition,” Cell, 65 175 –187 https://doi.org/10.1016/0092-8674(91)90418-X CELLB5 0092-8674 (1991). Google Scholar

120. 

B. Malnic et al., “Combinatorial receptor codes for odors,” Cell, 96 713 –723 https://doi.org/10.1016/S0092-8674(00)80581-4 CELLB5 0092-8674 (1999). Google Scholar

121. 

P. Duchamp-Viret, M. A. Chaput and A. Duchamp, “Odor response properties of rat olfactory receptor neurons,” Science, 284 2171 –2174 https://doi.org/10.1126/science.284.5423.2171 SCIEAS 0036-8075 (1999). Google Scholar

122. 

M. Ma and G. M. Shepherd, “Functional mosaic organization of mouse olfactory receptor neurons,” Proc. Natl. Acad. Sci. U. S. A., 97 12869 –12874 https://doi.org/10.1073/pnas.220301797 (2000). Google Scholar

123. 

P. Mombaerts et al., “Visualizing an olfactory sensory map,” Cell, 87 675 –686 https://doi.org/10.1016/S0092-8674(00)81387-2 CELLB5 0092-8674 (1996). Google Scholar

124. 

H. Takeuchi and H. Sakano, “Neural map formation in the mouse olfactory system,” Cell Mol. Life Sci., 71 3049 –3057 https://doi.org/10.1007/s00018-014-1597-0 CMLSFI 1420-9071 (2014). Google Scholar

125. 

T. Bozza et al., “Odorant receptor expression defines functional units in the mouse olfactory system,” J. Neurosci., 22 3033 –3043 https://doi.org/10.1523/JNEUROSCI.22-08-03033.2002 JNRSDS 0270-6474 (2002). Google Scholar

126. 

P. Feinstein et al., “Axon guidance of mouse olfactory sensory neurons by odorant receptors and the beta2 adrenergic receptor,” Cell, 117 833 –846 https://doi.org/10.1016/j.cell.2004.05.013 CELLB5 0092-8674 (2004). Google Scholar

127. 

J. L. Price and T. P. Powell, “The mitral and short axon cells of the olfactory bulb,” J. Cell Sci., 7 631 –651 https://doi.org/10.1242/jcs.7.3.631 JNCSAI 0021-9533 (1970). Google Scholar

128. 

K. M. Igarashi et al., “Parallel mitral and tufted cell pathways route distinct odor information to different targets in the olfactory cortex,” J. Neurosci., 32 7970 –7985 https://doi.org/10.1523/JNEUROSCI.0154-12.2012 JNRSDS 0270-6474 (2012). Google Scholar

129. 

S. Nagayama, R. Homma and F. Imamura, “Neuronal organization of olfactory bulb circuits,” Front. Neural Circuits, 8 98 https://doi.org/10.3389/fncir.2014.00098 (2014). Google Scholar

130. 

S. Parrish-Aungst et al., “Quantitative analysis of neuronal diversity in the mouse olfactory bulb,” J. Comp. Neurol., 501 825 –836 https://doi.org/10.1002/cne.21205 JCNEAM 0021-9967 (2007). Google Scholar

131. 

M. Najac et al., “Monosynaptic and polysynaptic feed-forward inputs to mitral cells from olfactory sensory neurons,” J. Neurosci., 31 8722 –8729 https://doi.org/10.1523/JNEUROSCI.0527-11.2011 JNRSDS 0270-6474 (2011). Google Scholar

132. 

D. H. Gire et al., “Mitral cells in the olfactory bulb are mainly excited through a multistep signaling path,” J. Neurosci., 32 2964 –2975 https://doi.org/10.1523/JNEUROSCI.5580-11.2012 JNRSDS 0270-6474 (2012). Google Scholar

133. 

D. De Saint Jan et al., “External tufted cells drive the output of olfactory bulb glomeruli,” J. Neurosci., 29 2043 –2052 https://doi.org/10.1523/JNEUROSCI.5317-08.2009 JNRSDS 0270-6474 (2009). Google Scholar

134. 

J. P. McGann, “Presynaptic inhibition of olfactory sensory neurons: new mechanisms and potential functions,” Chem. Sens., 38 459 –474 https://doi.org/10.1093/chemse/bjt018 (2013). Google Scholar

135. 

J. L. Aungst et al., “Centre-surround inhibition among olfactory bulb glomeruli,” Nature, 426 623 –629 https://doi.org/10.1038/nature02185 (2003). Google Scholar

136. 

W. Rall et al., “Dendrodendritic synaptic pathway for inhibition in the olfactory bulb,” Exp. Neurol., 14 44 –56 https://doi.org/10.1016/0014-4886(66)90023-9 EXNEAC 0014-4886 (1966). Google Scholar

137. 

G. M. Shepherd, W. R. Chen and C. A. Greer, Olfactory Bulb. In the Synaptic Organization of the Brain, 165 –216 Oxford University Press, New York (2004). Google Scholar

138. 

M. Wachowiak et al., “Inhibition [corrected] of olfactory receptor neuron input to olfactory bulb glomeruli mediated by suppression of presynaptic calcium influx,” J. Neurophysiol., 94 2700 –2712 https://doi.org/10.1152/jn.00286.2005 JONEA4 0022-3077 (2005). Google Scholar

139. 

D. A. Berkowicz and P. Q. Trombley,, “Dopaminergic modulation at the olfactory nerve synapse,”,” Brain Res., 855 90 –99 https://doi.org/10.1016/S0006-8993(99)02342-2 BRREAP 0006-8993 (2000). Google Scholar

140. 

A. M. Boyd et al., “Cortical feedback control of olfactory bulb circuits,” Neuron, 76 1161 –1174 https://doi.org/10.1016/j.neuron.2012.10.020 NERNET 0896-6273 (2012). Google Scholar

141. 

V. Kapoor et al., “Activation of raphe nuclei triggers rapid and distinct effects on parallel olfactory bulb output channels,” Nat. Neurosci., 19 271 –282 https://doi.org/10.1038/nn.4219 NANEFN 1097-6256 (2016). Google Scholar

142. 

F. Markopoulos et al., “Functional properties of cortical feedback projections to the olfactory bulb,” Neuron, 76 1175 –1188 https://doi.org/10.1016/j.neuron.2012.10.028 NERNET 0896-6273 (2012). Google Scholar

143. 

G. C. Petzold, A. Hagiwara and V. N. Murthy, “Serotonergic modulation of odor input to the mammalian olfactory bulb,” Nat. Neurosci., 12 784 –791 https://doi.org/10.1038/nn.2335 NANEFN 1097-6256 (2009). Google Scholar

144. 

M. Rothermel et al., “Cholinergic inputs from Basal forebrain add an excitatory bias to odor coding in the olfactory bulb,” J. Neurosci., 34 4654 –4664 https://doi.org/10.1523/JNEUROSCI.5026-13.2014 JNRSDS 0270-6474 (2014). Google Scholar

145. 

M. Rothermel and M. Wachowiak, “Functional imaging of cortical feedback projections to the olfactory bulb,” Front. Neural Circuits, 8 73 https://doi.org/10.3389/fncir.2014.00073 (2014). Google Scholar

146. 

C. Martelli and D. A. Storace, “Stimulus driven functional transformations in the early olfactory system,” Front. Cell. Neurosci., 15 684742 https://doi.org/10.3389/fncel.2021.684742 (2021). Google Scholar

147. 

A. K. Moran, T. P. Eiting and M. Wachowiak, “Dynamics of glutamatergic drive underlie diverse responses of olfactory bulb outputs,” eNeuro, 8 ENEURO.0110-21.2021 https://doi.org/10.1523/ENEURO.0110-21.2021 (2021). Google Scholar

148. 

D. A. Storace and L. B. Cohen, “Measuring the olfactory bulb input-output transformation reveals a contribution to the perception of odorant concentration invariance,” Nat. Commun., 8 81 https://doi.org/10.1038/s41467-017-00036-2 NCAOBW 2041-1723 (2017). Google Scholar

149. 

D. A. Storace and L. B. Cohen, “The mammalian olfactory bulb contributes to the adaptation of odor responses: a second perceptual computation carried out by the bulb,” eNeuro, 8 ENEURO.0322-21.2021 https://doi.org/10.1523/ENEURO.0322-21.2021 (2021). Google Scholar

150. 

D. A. Storace, L. B. Cohen and Y. Choi, “Using genetically encoded voltage indicators (GEVIs) to study the input-output transformation of the mammalian olfactory bulb,” Front. Cell Neurosci., 13 342 https://doi.org/10.3389/fncel.2019.00342 (2019). Google Scholar

151. 

R. Vincis et al., “Dense representation of natural odorants in the mouse olfactory bulb,” Nat. Neurosci., 15 537 –539 https://doi.org/10.1038/nn.3057 NANEFN 1097-6256 (2012). Google Scholar

152. 

M. N. Economo, K. R. Hansen and M. Wachowiak, “Control of mitral/tufted cell output by selective inhibition among olfactory bulb glomeruli,” Neuron, 91 397 –411 https://doi.org/10.1016/j.neuron.2016.06.001 NERNET 0896-6273 (2016). Google Scholar

153. 

L. M. Kay and G. Laurent, “Odor- and context-dependent modulation of mitral cell activity in behaving rats,” Nat. Neurosci., 2 1003 –1009 https://doi.org/10.1038/14801 NANEFN 1097-6256 (1999). Google Scholar

154. 

D. Rinberg, A. Koulakov and A. Gelperin, “Sparse odor coding in awake behaving mice,” J. Neurosci., 26 8857 –8865 https://doi.org/10.1523/JNEUROSCI.0884-06.2006 JNRSDS 0270-6474 (2006). Google Scholar

155. 

R. Shusterman et al., “Precise olfactory responses tile the sniff cycle,” Nat. Neurosci., 14 1039 –1044 https://doi.org/10.1038/nn.2877 NANEFN 1097-6256 (2011). Google Scholar

156. 

J. Niessing and R. W. Friedrich, “Olfactory pattern classification by discrete neuronal network states,” Nature, 465 47 –52 https://doi.org/10.1038/nature08961 (2010). Google Scholar

157. 

B. D. Rubin and L. C. Katz, “Optical imaging of odorant representations in the mammalian olfactory bulb,” Neuron, 23 499 –511 https://doi.org/10.1016/S0896-6273(00)80803-X NERNET 0896-6273 (1999). Google Scholar

158. 

R. Vincis et al., “Sensory-evoked intrinsic imaging signals in the olfactory bulb are independent of neurovascular coupling,” Cell Rep., 12 313 –325 https://doi.org/10.1016/j.celrep.2015.06.016 (2015). Google Scholar

159. 

M. Wachowiak and L. B. Cohen, “Correspondence between odorant-evoked patterns of receptor neuron input and intrinsic optical signals in the mouse olfactory bulb,” J. Neurophysiol., 89 1623 –1639 https://doi.org/10.1152/jn.00747.2002 JONEA4 0022-3077 (2003). Google Scholar

160. 

E. R. Soucy et al., “Precision and diversity in an odor map on the olfactory bulb,” Nat. Neurosci., 12 210 –220 https://doi.org/10.1038/nn.2262 NANEFN 1097-6256 (2009). Google Scholar

161. 

G. G. Blasdel and G. Salama, “Voltage-sensitive dyes reveal a modular organization in monkey striate cortex,” Nature, 321 579 –585 https://doi.org/10.1038/321579a0 (1986). Google Scholar

162. 

A. Grinvald et al., “Functional architecture of cortex revealed by optical imaging of intrinsic signals,” Nature, 324 361 –364 https://doi.org/10.1038/324361a0 (1986). Google Scholar

163. 

R. Chery et al., “Imaging odor-evoked activities in the mouse olfactory bulb using optical reflectance and autofluorescence signals,” J. Vis. Exp., 56 e3336 https://doi.org/10.3791/3336 (2011). Google Scholar

164. 

C. R. Uytingco, A. C. Puche and S. D. Munger, “Using intrinsic flavoprotein and NAD(P)H imaging to map functional circuitry in the main olfactory bulb,” PLoS One, 11 e0165342 https://doi.org/10.1371/journal.pone.0165342 POLNCL 1932-6203 (2016). Google Scholar

165. 

H. Spors and A. Grinvald, “Spatio-temporal dynamics of odor representations in the mammalian olfactory bulb,” Neuron, 34 301 –315 https://doi.org/10.1016/S0896-6273(02)00644-X NERNET 0896-6273 (2002). Google Scholar

166. 

R. W. Friedrich, “Calcium imaging in the intact olfactory system of zebrafish and mouse,” Cold Spring Harb. Protoc., 2014 310 –316 https://doi.org/10.1101/pdb.prot081166 (2014). Google Scholar

167. 

T. Bozza et al., “In vivo imaging of neuronal activity by targeted expression of a genetically encoded probe in the mouse,” Neuron, 42 9 –21 https://doi.org/10.1016/S0896-6273(04)00144-8 NERNET 0896-6273 (2004). Google Scholar

168. 

M. D. Kass et al., “Changes in olfactory sensory neuron physiology and olfactory perceptual learning after odorant exposure in adult mice,” Chem. Sens., 41 123 –133 https://doi.org/10.1093/chemse/bjv065 (2016). Google Scholar

169. 

J. P. McGann et al., “Odorant representations are modulated by intra- but not interglomerular presynaptic inhibition of olfactory sensory neurons,” Neuron, 48 1039 –1053 https://doi.org/10.1016/j.neuron.2005.10.031 NERNET 0896-6273 (2005). Google Scholar

170. 

A. Banerjee et al., “An interglomerular circuit gates glomerular output and implements gain control in the mouse olfactory bulb,” Neuron, 87 193 –207 https://doi.org/10.1016/j.neuron.2015.06.019 NERNET 0896-6273 (2015). Google Scholar

171. 

D. G. Blauvelt et al., “Distinct spatiotemporal activity in principal neurons of the mouse olfactory bulb in anesthetized and awake states,” Front. Neural Circuits, 7 46 https://doi.org/10.3389/fncir.2013.00046 (2013). Google Scholar

172. 

M. L. Fletcher et al., “Optical imaging of postsynaptic odor representation in the glomerular layer of the mouse olfactory bulb,” J. Neurophysiol., 102 817 –830 https://doi.org/10.1152/jn.00020.2009 JONEA4 0022-3077 (2009). Google Scholar

173. 

M. C. Ogg et al., “Olfactory bulb acetylcholine release dishabituates odor responses and reinstates odor investigation,” Nat. Commun., 9 1868 https://doi.org/10.1038/s41467-018-04371-w NCAOBW 2041-1723 (2018). Google Scholar

174. 

M. L. Fletcher, “Analytical processing of binary mixture information by olfactory bulb glomeruli,” PLoS One, 6 e29360 https://doi.org/10.1371/journal.pone.0029360 POLNCL 1932-6203 (2011). Google Scholar

175. 

S. Inagaki et al., “Widespread inhibition, antagonism, and synergy in mouse olfactory sensory neurons in vivo,” Cell Rep., 31 107814 https://doi.org/10.1016/j.celrep.2020.107814 (2020). Google Scholar

176. 

M. D. Kass et al., “Fear learning enhances neural responses to threat-predictive sensory stimuli,” Science, 342 1389 –1392 https://doi.org/10.1126/science.1244916 SCIEAS 0036-8075 (2013). Google Scholar

177. 

M. C. Ogg, M. Bendahamane and M. L. Fletcher, “Habituation of glomerular responses in the olfactory bulb following prolonged odor stimulation reflects reduced peripheral input,” Front. Mol. Neurosci., 8 53 https://doi.org/10.3389/fnmol.2015.00053 (2015). Google Scholar

178. 

J. M. Ross and M. L. Fletcher, “Learning-dependent and -independent enhancement of mitral/tufted cell glomerular odor responses following olfactory fear conditioning in awake mice,” J. Neurosci., 38 4623 –4640 https://doi.org/10.1523/JNEUROSCI.3559-17.2018 JNRSDS 0270-6474 (2018). Google Scholar

179. 

J. Platisa et al., “Voltage imaging in the olfactory bulb using transgenic mouse lines expressing the genetically encoded voltage indicator ArcLight,” Sci. Rep., 12 1875 https://doi.org/10.1038/s41598-021-04482-3 SRCEC3 2045-2322 (2022). Google Scholar

180. 

M. A. Passini et al., “Intraventricular brain injection of adeno-associated virus type 1 (AAV1) in neonatal mice results in complementary patterns of neuronal transduction to AAV2 and total long-term correction of storage lesions in the brains of beta-glucuronidase-deficient mice,” J. Virol., 77 7034 –7040 https://doi.org/10.1128/JVI.77.12.7034-7040.2003 JOVIAM 0022-538X (2003). Google Scholar

181. 

D. Vucinic, L. B. Cohen and E. K. Kosmidis, “Interglomerular center-surround inhibition shapes odorant-evoked input to the mouse olfactory bulb in vivo,” J. Neurophysiol., 95 1881 –1887 https://doi.org/10.1152/jn.00918.2005 JONEA4 0022-3077 (2006). Google Scholar

182. 

H. Spors et al., “Temporal dynamics and latency patterns of receptor neuron input to the olfactory bulb,” J. Neurosci., 26 1247 –1259 https://doi.org/10.1523/JNEUROSCI.3100-05.2006 JNRSDS 0270-6474 (2006). Google Scholar

183. 

E. C. Sobel and D. W. Tank, “Timing of odor stimulation does not alter patterning of olfactory bulb unit activity in freely breathing rats,” J. Neurophysiol., 69 1331 –1337 https://doi.org/10.1152/jn.1993.69.4.1331 JONEA4 0022-3077 (1993). Google Scholar

184. 

K. B. Doving, “Response properties of neurones in the rat olfactory bulb to various parameters of odour stimulation,” Acta Physiol. Scand., 130 285 –298 https://doi.org/10.1111/j.1748-1716.1987.tb08139.x (1987). Google Scholar

185. 

J. Cang and J. S. Isaacson, “In vivo whole-cell recording of odor-evoked synaptic transmission in the rat olfactory bulb,” J. Neurosci., 23 4108 –4116 https://doi.org/10.1523/JNEUROSCI.23-10-04108.2003 JNRSDS 0270-6474 (2003). Google Scholar

186. 

M. Mayford et al., “Control of memory formation through regulated expression of a CaMKII transgene,” Science, 274 1678 –1683 https://doi.org/10.1126/science.274.5293.1678 SCIEAS 0036-8075 (1996). Google Scholar

187. 

J. Li et al., “Odorant receptor gene choice is reset by nuclear transfer from mouse olfactory sensory neurons,” Nature, 428 393 –399 https://doi.org/10.1038/nature02433 (2004). Google Scholar

188. 

Y. Nagai, H. Sano and M. Yokoi, “Transgenic expression of Cre recombinase in mitral/tufted cells of the olfactory bulb,” Genesis, 43 12 –16 https://doi.org/10.1002/gene.20146 (2005). Google Scholar

189. 

M. Kohwi et al., “A subpopulation of olfactory bulb GABAergic interneurons is derived from Emx1- and Dlx5/6-expressing progenitors,” J. Neurosci., 27 6878 –6891 https://doi.org/10.1523/JNEUROSCI.0254-07.2007 JNRSDS 0270-6474 (2007). Google Scholar

190. 

S. Firestein, C. Picco and A. Menini, “The relation between stimulus and response in olfactory receptor cells of the tiger salamander,” J. Physiol., 468 1 –10 https://doi.org/10.1113/jphysiol.1993.sp019756 JPHYA7 0022-3751 (1993). Google Scholar

191. 

T. Kurahashi and A. Menini, “Mechanism of odorant adaptation in the olfactory receptor cell,” Nature, 385 725 –729 https://doi.org/10.1038/385725a0 (1997). Google Scholar

192. 

A. Menini, C. Picco and S. Firestein, “Quantal-like current fluctuations induced by odorants in olfactory receptor cells,” Nature, 373 435 –437 https://doi.org/10.1038/373435a0 (1995). Google Scholar

193. 

J. Reisert and H. R. Matthews, “Adaptation of the odour-induced response in frog olfactory receptor cells,” J. Physiol., 519 (3), 801 –813 https://doi.org/10.1111/j.1469-7793.1999.0801n.x JPHYA7 0022-3751 (1999). Google Scholar

194. 

J. Reisert and H. R. Matthews, “Response properties of isolated mouse olfactory receptor cells,” J. Physiol., 530 113 –122 https://doi.org/10.1111/j.1469-7793.2001.0113m.x JPHYA7 0022-3751 (2001). Google Scholar

195. 

J. Reisert and H. R. Matthews, “Responses to prolonged odour stimulation in frog olfactory receptor cells,” J. Physiol., 534 179 –191 https://doi.org/10.1111/j.1469-7793.2001.t01-1-00179.x JPHYA7 0022-3751 (2001). Google Scholar

196. 

N. Uchida and Z. F. Mainen, “Odor concentration invariance by chemical ratio coding,” Front. Syst. Neurosci., 1 3 https://doi.org/10.3389/neuro.06.003.2007 (2007). Google Scholar

197. 

R. Gross-Isseroff and D. Lancet, “Concentration-dependent changes of perceived odor quality,” Chem. Sens., 13 191 –204 https://doi.org/10.1093/chemse/13.2.191 (1988). Google Scholar

198. 

D. Rokni et al., “An olfactory cocktail party: figure-ground segregation of odorants in rodents,” Nat. Neurosci., 17 1225 –1232 https://doi.org/10.1038/nn.3775 NANEFN 1097-6256 (2014). Google Scholar

199. 

K. Asahina et al., “A circuit supporting concentration-invariant odor perception in Drosophila,” J. Biol., 8 9 https://doi.org/10.1186/jbiol108 (2009). Google Scholar

200. 

R. Homma et al., “Perceptual stability during dramatic changes in olfactory bulb activation maps and dramatic declines in activation amplitudes,” Eur. J. Neurosci., 29 1027 –1034 https://doi.org/10.1111/j.1460-9568.2009.06644.x EJONEI 0953-816X (2009). Google Scholar

201. 

T. A. Cleland et al., “Relational representation in the olfactory system,” Proc. Natl. Acad. Sci. U. S. A., 104 1953 –1958 https://doi.org/10.1073/pnas.0608564104 (2007). Google Scholar

202. 

T. A. Cleland, “Early transformations in odor representation,” Trends Neurosci., 33 130 –139 https://doi.org/10.1016/j.tins.2009.12.004 TNSCDR 0166-2236 (2010). Google Scholar

203. 

M. Meredith, “Patterned response to odor in mammalian olfactory bulb: the influence of intensity,” J. Neurophysiol., 56 572 –597 https://doi.org/10.1152/jn.1986.56.3.572 JONEA4 0022-3077 (1986). Google Scholar

204. 

H. S. Orbach and L. B. Cohen, “Optical monitoring of activity from many areas of the in vitro and in vivo salamander olfactory bulb: a new method for studying functional organization in the vertebrate central nervous system,” J. Neurosci., 3 2251 –2262 https://doi.org/10.1523/JNEUROSCI.03-11-02251.1983 JNRSDS 0270-6474 (1983). Google Scholar

205. 

J. A. London, D. Zecević and L. B. Cohen, “Simultaneous optical recording of activity from many neurons during feeding in Navanax,” J. Neurosci., 7 649 –661 https://doi.org/10.1523/JNEUROSCI.07-03-00649.1987 JNRSDS 0270-6474 (1987). Google Scholar

206. 

S. Antić and D. Zecević, “Optical signals from neurons with internally applied voltage-sensitive dyes,” J. Neurosci., 15 1392 –1405 https://doi.org/10.1523/JNEUROSCI.15-02-01392.1995 JNRSDS 0270-6474 (1995). Google Scholar

207. 

M. Djurisic et al., “Voltage imaging from dendrites of mitral cells: EPSP attenuation and spike trigger zones,” J. Neurosci., 24 6703 –6714 https://doi.org/10.1523/JNEUROSCI.0307-04.2004 JNRSDS 0270-6474 (2004). Google Scholar

208. 

M. A. Popovic et al., “Electrical behaviour of dendritic spines as revealed by voltage imaging,” Nat. Commun., 6 8436 https://doi.org/10.1038/ncomms9436 NCAOBW 2041-1723 (2015). Google Scholar

209. 

P. Chakrabarty et al., “Capsid serotype and timing of injection determines AAV transduction in the neonatal mice brain,” PLoS One, 8 e67680 https://doi.org/10.1371/journal.pone.0067680 POLNCL 1932-6203 (2013). Google Scholar

210. 

Y. Adam et al., “Voltage imaging and optogenetics reveal behaviour-dependent changes in hippocampal dynamics,” Nature, 569 413 –417 https://doi.org/10.1038/s41586-019-1166-7 (2019). Google Scholar

211. 

W. Denk, J. H. Strickler and W. W. Webb, “Two-photon laser scanning fluorescence microscopy,” Science, 248 73 –76 https://doi.org/10.1126/science.2321027 SCIEAS 0036-8075 (1990). Google Scholar

212. 

R. D. Frostig, In vivo optical imaging of brain function, CRC Press( (2009). Google Scholar

213. 

W. Akemann et al., “Two-photon voltage imaging using a genetically encoded voltage indicator,” Sci. Rep., 3 2231 https://doi.org/10.1038/srep02231 SRCEC3 2045-2322 (2013). Google Scholar

214. 

K. F. Ahrens et al., “Two-photon scanning microscopy of in vivo sensory responses of cortical neurons genetically encoded with a fluorescent voltage sensor in rat,” Front. Neural Circuits, 6 15 https://doi.org/10.3389/fncir.2012.00015 (2012). Google Scholar

215. 

J. Platisa and V. A. Pieribone, “Genetically encoded fluorescent voltage indicators: are we there yet?,” Curr. Opin. Neurobiol., 50 146 –153 https://doi.org/10.1016/j.conb.2018.02.006 COPUEN 0959-4388 (2018). Google Scholar

216. 

R. Homma et al., “Narrowly confined and glomerulus-specific onset latencies of odor-evoked calcium transients in the juxtaglomerular cells of the mouse main olfactory bulb,” eNeuro, 6 ENEURO.0387-18.2019 https://doi.org/10.1523/ENEURO.0387-18.2019 (2019). Google Scholar

217. 

H. Dana et al., “Sensitive red protein calcium indicators for imaging neural activity,” Elife, 5 e12727 https://doi.org/10.7554/eLife.12727 (2016). Google Scholar

Biography

Lee Min Leong received her PhD in biomedical science and technology from the Korea Institute of Science and Technology in Seoul, South Korea, where she worked on GEVI development and understanding the mechanism of GEVIs in the laboratory of Bradley J. Baker. She is currently a postdoctoral fellow in the laboratory of Douglas A. Storace at Florida State University using voltage and calcium sensors in the olfactory system.

Douglas A. Storace received his PhD in behavioral neuroscience from the University of Connecticut in Storrs, Connecticut, before completing a postdoctoral fellowship in the laboratory of Lawrence B. Cohen. He is currently an assistant professor at Florida State University in the Department of Biological Science. His research group uses in vivo widefield and 2-photon microscopy to understand sensory coding in the olfactory system.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Lee Min Leong and Douglas A. Storace "Imaging different cell populations in the mouse olfactory bulb using the genetically encoded voltage indicator ArcLight," Neurophotonics 11(3), 033402 (17 January 2024). https://doi.org/10.1117/1.NPh.11.3.033402
Received: 1 September 2023; Accepted: 14 December 2023; Published: 17 January 2024
Advertisement
Advertisement
KEYWORDS
Fluorescence

Calcium

Neurons

In vivo imaging

Action potentials

Brain

Dyes

Back to Top