Open Access
30 November 2023 Advances in cellular resolution microscopy for brain imaging in rats
Su Jin Kim, Rifqi O. Affan, Hadas Frostig, Benjamin B. Scott, Andrew S. Alexander
Author Affiliations +
Abstract

Rats are used in neuroscience research because of their physiological similarities with humans and accessibility as model organisms, trainability, and behavioral repertoire. In particular, rats perform a wide range of sophisticated social, cognitive, motor, and learning behaviors within the contexts of both naturalistic and laboratory environments. Further progress in neuroscience can be facilitated by using advanced imaging methods to measure the complex neural and physiological processes during behavior in rats. However, compared with the mouse, the rat nervous system offers a set of challenges, such as larger brain size, decreased neuron density, and difficulty with head restraint. Here, we review recent advances in in vivo imaging techniques in rats with a special focus on open-source solutions for calcium imaging. Finally, we provide suggestions for both users and developers of in vivo imaging systems for rats.

1.

Introduction

Advances in genetically encoded sensors provide increased sensitivity, cell type specificity, and the ability to record a variety of signals from intracellular calcium1 and membrane voltage,2 to neurotransmitter release such as dopamine.3,4 New microscopes have been developed to image across larger areas, with greater resolution, increased depth, and enhanced portability.57 These methods are being increasingly paired with sophisticated analytical techniques, which have opened new avenues within theoretical neuroscience.810

The development of in vivo cellular resolution imaging technologies, and calcium imaging in particular, has been one of the modern success stories in systems neuroscience.11 Over the past 60 years, these tools have been applied to a variety of model organisms [Fig. 1(a)]. However, in the last 15 years, the mouse has emerged as a leading model for in vivo cellular resolution imaging. This is likely due to the confluence of genetic tools, such as transgenic mouse lines (e.g., Ref. 23), and methods that enable imaging during behavior, such as head-fixed virtual reality (VR16) and head mounted microscopes.17

Fig. 1

Tasks and behavioral control systems used in rats. (a) Number of papers on PubMed by year with the search term “calcium imaging” and either “rat,” “mouse,” “zebrafish,” “drosophila,” or “Caenorhabditis elegans” from 1964 to 2019. Key calcium imaging papers are denoted by a triangle and the citation: development of fura-2, a fluorescent dye to detect calcium,12 the first 2P microscope,13 development of an early genetically encoded calcium sensor,14 a treadmill system for in vivo imaging,15 the first VR system used with 2P imaging in mice,16 wearable epifluorescent microscope,17 development of GCaMP6,18 and development of the open-source miniscope.19 (b) Schematic of a tactile comparison task to measure parametric working memory (top), with rats (middle) and humans (bottom) performing the task.20 (c) Trajectory of a hide-and-seek task trial in rats, where the rat emerges from the start box and searches for the human experimenter.21 (d) A fully automated, live-in facility for rat behavioral training.22

NPh_10_4_044304_f001.png

While success of in vivo imaging technologies in mice has inspired the field, progress in other organisms, including rats, continues. Rats have historically been an important model for biomedical and neuroscience research (Refs. 24 and 25; see Table 1). Today they remain a leading model for studying neural dynamics during complex learned behaviors, such as navigation, decision making, and addiction. The behavioral advantages of this organism have motivated continued innovation in applying calcium imaging tools. Recent successes reflect this: new imaging technologies for rats include multiphoton microscopy using voluntary head restraint,34,35 open-source widefield microscopes for large field of view (FOV) recording,36,37 head mounted three photon (3P) microscopes,38 and transgenic rats expressing calcium indicators.34,37

Table 1

Pioneering discoveries in systems neuroscience using the rat model.

DiscoveryReference(s)
Adult neurogenesisAltman and Das26
Place cellsO’Keefe and Dostrovsky27
Head direction cellsTaube et al.28
First BOLD measurement with fMRIOgawa et al.29
Odorant receptor geneBuck and Axel30
Neural replayWilson and McNaughton31
In vivo 2P imagingDenk et al.32
Grid cellsHafting et al.33

Given the significance of the rat animal model in neuroscience and neuroimaging specifically, continued development of in vivo imaging tools in this species is warranted. This review will focus on specific opportunities and challenges posed by neuroimaging in the rat model, describe the technical solutions under development, and provide an outlook for technologies that may facilitate future imaging experiments.

2.

Opportunities in Rat Imaging

The rat model has advantages that motivate its continued use for studying the link between cellular dynamics and behavior. In this section, we provide an overview of these advantages and the experimental opportunities of the rat model system.

2.1.

Behavioral Repertoire

Rats distinguish themselves as model organisms because of their complex behavioral repertoire, adaptability, and the variety of tools to study both learned and natural behaviors. Rats can be trained on a wide range of tasks designed to characterize goal-directed behaviors and decision-making.3943 For example, rats can readily learn to perform parametric working memory tasks inspired by primate tasks20,44 [Fig. 1(b)] and can learn the representation of action-outcome associations in a multi-step planning tasks.45,46 Rats can learn behavioral paradigms originally developed for humans, facilitating comparative studies and translational research in neuropsychiatry.47,48 Rats are also social creatures,49 demonstrating pro-social behaviors in controlled laboratory environments,50,51 including empathy,52 cross species play21 [Fig. 1(c)], and collaborative group search.53

The wide range of behavioral features in rats contribute to their usefulness as a model organism for basic and translational neuroscience research. Unfortunately, direct, quantitative comparisons of behaviors between rats and other model organisms, in particular mice, is rarely performed, and this limitation is particularly acute in complex decision-making tasks, which are presently of great interest.54 Ethological behaviors are somewhat conserved; mice and rats have similar aggression, grooming, feeding, and reproductive behaviors.55 While the overall behavioral patterns are consistent between species, there are slight nuances to many of these innate behaviors (e.g., rats exhibit more complex grooming phases than mice). A quantitative comparison between rat and mouse behavior across a range of tasks would facilitate an unbiased assessment of the pros and cons of each species. In some cases, such as addiction, these side-by-side comparisons have been performed. For example, there is some indication that rats are a better model for studying alcohol relapse behaviors than mice.56

Numerous open-source tools and pipelines have been developed for behavioral training and measurement in rats. These include VR navigation systems5759 automated operant systems22,39 [Fig. 1(d)], touchscreen training,60 and voluntary head restraint.6163 Together, the availability of experimental and computational tools for behavioral research in rats provides frameworks for collecting and analyzing high-throughput data in a variety of laboratory settings, which can easily be paired with multimodal imaging approaches.64

2.2.

Body Size

Adult rats weigh hundreds of grams (250 to 350 g for a 10 week old male Long Evans)65 and have significant capacity for implantable and wearable devices. Rats can carry head mounted devices weighing 35 g while still displaying natural behaviors, such as rearing and rapid head orienting.37 This capacity reduces constraints on development allowing for microscopes with larger FOVs36,37 and or more complex optical components.66 Beyond the rat’s physical strength, the larger size and rectilinear shape of the skull provides ample “real estate” for device attachment.

Beyond the technological advantages that rats provide because of their physiology, rats can also act as a bridge to larger model organisms for neuroscientific research. As we describe below, the brains of larger animals pose challenges to imaging, which will require new imaging capabilities. Rats, with their relatively wide range of available transgenic lines and genetic tools, may provide a valuable test case for developing and expanding technology for other animals, such as ferrets, macaques, and marmosets.

3.

Challenges in Rat Imaging

3.1.

Head Restraint

Head restraint is widely used in neuroscience to stabilize the brain position relative to the imaging apparatus. Head restraint in rats can be accomplished through an acclimation process in which the duration of restraint is gradually increased.67 However, compared with mice, this approach is unreliable and more limited in rats—they show increased stress and diminished behavioral flexibility during head restraint.68 Consequently, forced head restraint is not frequently used in conjunction with complex cognitive task learning in rats. This has motivated the development of head-mounted microscopes and voluntary head-fixation (see Sec. 4).

3.2.

Decreased Neuronal Density

While being 8 to 10 times the body mass of mice, rats have three times the number of neurons; much of this increase in neurons is in the cerebellum, and the fraction of cortical neurons remains constant even as total brain size increases.6971 Mice have on average 78,672 neurons and 68,640 nonneuronal cells per milligram of cortical tissue, whereas rats have 41,092 neurons and 60,430 nonneuronal cells per milligram.69 In terms of density, rats have half the number of neurons per milligram of cerebral cortex compared with mice.70,72 Lower neuron densities will result in fewer imaged neurons when assuming the same FOV and signal-to-noise ratio (SNR). This challenge is not unique to rats—it is a challenge shared by many larger-brained animals, including several primate species.69,70,73

3.3.

Increased Cortical Thickness

The rat neocortex is thicker than the mouse neocortex; for example, the motor cortex of rats has an average thickness of 1.6 mm while in mice motor cortex has an average thickness of 1.0 mm.71 Since the scattering length of the rat cortex is similar to that of the mouse,32,7478 the excitation light penetrates to a comparable depth in both animals. Overall, this results in reduced optical access into deeper layers in the rat brain. In most cases, cell somas in layer 2/3 of rat neocortex, which ranges from 200 to 500  μm,79 can lie below the range of some head-mounted one-photon imaging systems17 and makes imaging of infragranular layers difficult. To surpass these limitations, researchers can implant microprisms, relay gradient index (GRIN) lenses, or use 3P microscopy, all three of which we discuss in more detail in the following section (see Sec. 4).

3.4.

Vascular Size and Branches

Rat brains have an increased number of capillary branches per unit volume and larger radii of vessels compared to mouse brains.80,81 This can lead to changes in the optical properties of tissue, such as increased absorption of light at different wavelengths due to hemoglobin.37,8284 In addition, these differences in vasculature can contribute to difficulties in surgery (such as increased bleeding) when compared to mice.

3.5.

Transgenesis

Tools for the production of transgenic rats are well developed and several lines of genetically modified rats that express calcium sensors have been produced (see Sec. 4). However, the costs, speed of generation, and number of off the shelf transgenic lines in mice greatly exceeds the rat model at present. The availability of transgenic lines is an important feature that should be considered when selecting a model organism for calcium imaging studies.

4.

Tools for Rat Imaging

Below, we highlight the recent applications of imaging tools and labeling techniques in rats.

4.1.

Transgenic Lines

Several useful transgenic lines for neuroscience and specifically in vivo imaging are available from several sources, including the Rat Resource and Research Center (RRRC) and the Rat Genome Database.85 Available lines include Cre driver lines for cell type specific expression (e.g. Refs. 86 to 88) and genetic models for human neuropsychiatric disorders, such as models of autism.89 The Rat Genome Database provides a valuable list of resources for the development of transgenic rats.90

Transgenic lines have also been developed that express genetically encoded calcium sensors for in vivo imaging34,37 [Figs. 2(a)2(c)]. These lines, created by Janelia Research Campus on the Long-Evans background, express the genetic calcium indicator GCaMP6f throughout large regions of the CNS, with different transgenic lines having clusters of expression in different areas.

Fig. 2

Labeling systems for rats. (a) Sagittal section of a Thy1 GCaMP6f-9 rat (from Ref. 37). (b), (c) 2P imaging of layer 2/3 of the cerebral cortex of a transgenic rat expressing GCAMP6f, where red pixels identify ROIs.37 (d) Calcium traces from the 17 ROIs in panel C at 30 Hz.37 (e) Epifluorescence image of a cranial window in a rat following serial viral injections with AAV9-GCaMP7f.67 (f) 2P imaging of a 500  μm×500  μm FOV from rat cortex injected with GCaMP7f.67 (g) Z-scored traces from the rat visual cortex for three cycles of presentation of a moving bar sweeping in the nasal-to-temporal direction at 0.24 Hz. Traces are colored and sorted by the corresponding cell’s phase at the stimulation frequency.67 (h) Confocal imaging of a coronal section of the rat hippocampus expressing Lck-GCaMP6f following in utero electroporation.91 (i) Mean calcium activity projection of a neuron expressing Lck-Gcamp6f following in utero electroporation and using 2P microscopy.91 (j) Calcium traces from the same cortical neuron, with colors corresponding to the dashed ROIs in panel (i).91

NPh_10_4_044304_f002.png

Sensor expression in at least two transgenic rat strains, Thy-1-GCaMP6f-7 and Thy-1-GCaMP6f line 8, is sufficient for cellular resolution imaging through either one or two-photon (2P) microscopes.34,37,92 However, the use of newer GCaMP variants delivered by adeno associated viral vectors (AAV) injection appears to provide improved SNR and action potential (AP) detection. For example, Chornyy et al.93 found that single AP detection was detected in 10.6% (48/450) of GCaMP6f-labeled neurons labeled in Thy-1-GCaMP6f animals, whereas it was detected in 85% (412/485) of jGCaMP7s-positive cells labeled with AAVs. These results indicate that new GCaMP variants and/or viral labeling may improve signal detection.

4.2.

Viral Vectors

At the time of writing, the majority of published studies involving imaging of genetically encoded sensors in rats express sensors using direct injection of AAVs35,36,63,9498 [Figs. 2(d) and 2(e)]. AAVs are favored due to the high-levels of expression that are difficult to obtain in transgenics99 and the availability of new genetically encoded sensors, which are being developed more rapidly than new transgenic lines. Direct injection of high-titer viral vectors into the rat CNS is widely used to achieve local expression of genetically encoded sensors. To achieve more widespread expression, several alternative approaches have been explored. One method is using serial injections, which has been demonstrated across the rat cortex. In this approach, a series of injections are performed at regular increments, tiling a larger volume. This approach aims to achieve a more uniform labeling over a larger volume than could be achieved by a single injection [Figs. 2(d) and 2(e)].67,88,100 Several groups have also reported widespread CNS infection in adults following systemic administration through intravenous,101103 intraventricular, and intrathecal injection.104,105 These techniques reduce the potential for damage to neural tissue following direct injections. The efficiency of these techniques is enhanced by the development of enhanced AAV capsids (such as PHP.eB), which yield improved gene transfer in rat CNS.104,106 While these approaches are intriguing, they have not been widely used in combination with in vivo functional imaging approaches in rats.

4.3.

In Utero Gene Delivery

Another method for gene delivery used in rats is via in utero electroporation, a method for transfecting neural tissue with plasmid DNA via injection into embryonic brains [Figs. 2(f) and 2(g)]. In utero electroporation enables widespread expression in neurons throughout the CNS.91,107111 In addition, in utero AAV injections can be used to achieve widespread cortical labeling in rats.112

A strength of in utero gene delivery is that it can be implemented during different stages of development to yield spatially specific expression within the neocortex without the need for laminar specific promoters. Moreover, the method can be optimized to produce widespread infection from a single injection. That said, gene delivery to the rat embryo requires specialized techniques and equipment, and there is some indication that introduction of foreign genetic material during development can produce an immune response that alters or even damages the brain.113

4.4.

Head-Mounted Microscopes Designed for Mice

Miniaturized head mounted epifluorescence microscopes allow recording of calcium dynamics in freely behaving animals.17 This approach bypasses the problem of head restraint and stabilization while achieving cellular resolution imaging.114 These microscopes have been widely used in mice, but several groups have applied these miniature microscopes in rats.92,9498,115 However, performance in these scopes is often optimized for mice. For example, early generations of UCLA microscopes have an FOV of 1  mm2. Next generation miniscopes designed with rats in mind have a larger FOV to account for the decreased cell density in the species (discussed in greater detail below).

The size and strength of the rat can create issues for the physical stability of head-mounted microscopes. Open-source systems, such as headcap covers, have been developed to protect and stabilize the scope.116 A headcap system for protecting the microscope also permits a solution for reducing movement-related torque on the microscope from the tethering cable. Once implanted, an anchoring point on the headcap offset from the microscope can be used to fix the tether to the headcap and thus reduce force transferred at the connection point with the microscope.

4.5.

Microprisms

As discussed, rat cortex is thicker relative to mouse cortex and this increased depth increases light scattering, decreases SNR, and prevents optical access to deep layers. One way to bypass these issues is to image through microprisms implanted directly into neural tissue as previously reported in mice.117,118 Recently, Alexander et al.92 successfully paired microprisms with head-mounted one-photon microscopes to image large populations of neurons in rat neocortex (Fig. 3). In this preparation, a 1  mm2 microprism attached to a relay lens was positioned near neurons expressing GCaMP6f to create an FOV perpendicular to the dorsal surface of the brain spanning multiple cortical layers [Figs. 3(a)3(c)]. A baseplate was attached to the skull above the microprism, which allowed a head-mounted microscope to be mounted [Fig. 3(d)]. Using this preparation, it was possible to simultaneously monitor calcium dynamics of hundreds of neurons with robust SNRs in Thy1-GCaMP6f transgenic rats performing track running or free exploration [Figs. 3(e)3(k)]. Well known spatial coding properties of the retrosplenial cortex (RSC) were replicated using this method in rats including trajectory-dependent coding [Fig. 3(h)] and coding for environmental boundaries in egocentric coordinates [Fig. 3(k)].

Fig. 3

Calcium imaging in transgenic rats through implanted microsprisms. (a) Position of the implanted microprism for imaging in the rat RSC relative to the rat head. (b) Schematic of the implantation location in a sagittal section. (c) Schematic of the prism imaging approach. (d) Image of an implanted rat wearing a head mounted one-photon camera in an operant chamber. (e) Maximum intensity projection from the imaging FOV. (f) Example time traces from selected ROIs from E showing fluorescence transients during an operant-based task. (g) Deconvolved Ca2+ traces from 30 simultaneously recorded RSC neurons. (h) Six RSC neurons, recorded using this preparation, exhibit differential activation for different trajectories on a delayed alternation spatial working memory task on a T-maze. Gray lines represent trajectory on track, split into leftward and rightward trials. Colored dots indicate animal position and head direction at the time of a calcium transient. Color indicates head direction according to legend on top right. (i) Number of cells per session. (j) Distribution of mean transient rate from a single recording. (k) Simultaneous recording of six RSC neurons with egocentric boundary vector responsivity. (Left) Trajectory plot with animal path in gray and spike locations indicated in colored circles where color is animal heading orientation in the environment. (Middle) Two-dimensional ratemap of “spiking” activity. (Right) Egocentric boundary ratemap showing position of boundaries at time of calcium transient. F, front; B, behind; R, right; L, left. All plots are maximum normalized (blue = zero activity, yellow = maximal).

NPh_10_4_044304_f003.png

4.6.

Head-Mounted 1P Microscopes Designed for Rats

Head-mounted widefield microscopes with larger FOVs have been developed for rats (Fig. 4). Larger FOVs enable the monitoring of larger populations of neurons and permit the examination of cross-regional dynamics not afforded by a smaller FOV targeting a single brain region. Previously, researchers developed cScope, a head mounted widefield macroscope to access FOVs up to 8  mm2 [Figs. 4(a)4(c)].37 cScope uses a hemodynamic illumination collar with green LEDs for reflectance illumination of cortical intrinsic signal and a blue LED for fluorescence imaging. Recordings using cScope have similar performance compared to conventional widefield epifluorescence microscopes, with imaging frame speed up to 30 Hz. However, the authors did not report cellular resolution calcium dynamics or whether this fluorescence signal originates from soma or neuropil.

Fig. 4

Head mounted widefield microscopes designed for rats. (a) Schematic of a rat wearing cScope, a head-mounted widefield macroscope.37 (b) Image of the FOV in a rat implanted with cScope. (c) Left: cScope fluorescence image, with colored dots indicating the location of the pixels that contribute to the responses on the right. Right: Flash response dynamics of the corresponding single pixel ROIs. (d) Picture of a rat wearing MiniLFOV.36 (e) Maximum projection of a motion-corrected recording session. Scale bar: 500  μm. (f) Left: Map within panel (e) of 59 cells. Scale bar: 100  μm. Right: Calcium traces from a subset of 15 cells within panel (f) across 6 min.

NPh_10_4_044304_f004.png

A recent implementation of the UCLA Miniscope, Miniscope-LFOV, was developed for rats [Figs. 4(d)4(f)].36 This system is a one-photon microscope, which has two electrically adjustable working distance (±100  μm) configurations that allow for cortical imaging via a cranial window and deep brain imaging via a relay GRIN lens. It has a 3.6  mm×2.7  mm FOV, with one FOV in CA1 revealing 1357 cells.36 The SNR in this microscope is considerably higher when compared with the performance of previous Miniscope iterations, attributable to newer and more sensitive detection systems in Miniscope-LFOV compared to its Miniscope predecessors. Recently published work details a system for online data pre-processing with Miniscope-LFOV,119 enabling researchers to perform motion correction, calcium trace extraction, and recognize neural patterns, which are correlated to behavior.

4.7.

Head-Mounted Multiphoton Microscopes Made for Rats

The carrying capacity of rats has facilitated the development of advanced head-mounted microscopes, such as multiphoton microscopes. The first head-mounted 2P microscope was developed for rats in the early 2000s, by Helmchen et al. [Figs. 5(a)5(e)].120 This microscope was 25 g in weight and 7.5 cm in height. Scanning was achieved by a fiber tip that resonated to form a Lissajous pattern. More recent iterations allow for increased performance, including raster scanning, and provide optical access to deeper areas with cellular resolution imaging in behaving rats [Figs. 5(f)5(h)].66 Today head mounted 3P microscopes for rats have cellular resolution as deep as 1.1 mm with a 150  μm square FOV [Figs. 5(i)5(k)]38 and more recently adapted to mice.121

Fig. 5

Head-mounted multiphoton microscopes used in rats. (a) Diagram of the light path and setup of the first head-mounted 2P microscope.120 (b) Schematic of the internal components in the fiberscope design. (c) Images of somatosensory cortex L2/3 neurons filled with calcium green-1. (d) Zoomed in image of a different dendrite from in somatosensory cortex L2/3. (e) Example calcium green-1 fluorescence trace along a dendritic process following current injection at 1 s intervals, with 10 ms resolution. (f) Picture of a rat wearing a head-mounted 2P microscope.66 (g) Camera image of the primary visual area, with the 2P imaging sites identified with the red dashed line. (h) Left: Two color 2P imaging of primary visual cortex using sulforhodamine 101 and OGB1-AM. Right: calcium time courses of the soma of three neurons (colored circles in the left panel) across 30 s. (i) Image of a 120 g rat wearing a head mounted three-photon microscope.38 (j) Histological section of GCaMP6s-labeled neurons in posterior parietal rat cortex, with the yellow dotted box showing the attainable imaging depth (1120  μm). (k) Left: Labeled neurons at 1120  μm depth below the cortical surface. Right: Example spontaneous calcium kinetics from FOV on left.

NPh_10_4_044304_f005.png

Like their tabletop counterparts, head-mounted multiphoton microscopes have several key features that facilitate calcium imaging in vivo in larger brained mammals, such as rats. The longer excitation wavelengths allow for less scattering in tissue and greater power delivery at depth. The non-linear properties of excitation provide optical sectioning and a reduction in out-of-focus excitation from fluorescence contamination from sources above and below the imaging plane.13,122 Multiphoton imaging can improve the ability to resolve cellular structures like axonal projections and dendrites in scattering tissue and can reduce contamination from the neuropil in vivo.123 However, head-mounted multiphoton microscopes are still outperformed by table top microscopes, including both commercial and custom systems, due to fewer space and weight constraints in the tabletop environment. Therefore, in order to combine the power to table top scopes with automated behavioral training systems, voluntary head restraint tools have been developed.

4.8.

Voluntary Head-Restraint

Voluntary head-restraint is a system in which trained rodents submit to periods of mechanical head restraint for reward (Fig. 6). Initially developed for rats for repeatable presentation of visual stimuli,61,62 demonstrations that computer controlled training systems for precise positioning and stability catalyzed renewed interest in voluntary head restraint.63,126 Work in rats inspired researchers to develop automated behavioral systems using voluntary head-fixation in mice.127130 These head fixation systems have been designed for mechanical stability and repositioning within several microns and to be used together with widefield imaging or optogenetics.

Fig. 6

Principles of voluntary head restraint. (a)–(c) A rat voluntarily head restraining across three stages: pre insertion, positioning of the head clamp and fixation, and release.63 (d) The principles of kinematic coupling. Objects can be exactly constrained with stable points equal to the degrees of freedom the object has, or over-constrained such that there are multiple stable points possible. Kinematic coupling enables high degrees of repeatability and accuracy by exactly constraining objects.124 (e) Toy model of a vee groove kinematic clamp.124 (f) Diagram of the degrees of freedom constrained in a vee groove kinematic clamp.125 (g) Behavioral paradigm schematic where rats are trained to voluntarily head restrain during an evidence accumulation task.42 (h) 2P imaging of GCaMP3-labeled cortical neurons across several voluntary head-restraint trials. Top panels show V1 without motion correction. Bottom panel shows fluorescence transients from the selected neuron (indicated by the white arrow). On each trial, a visual stimulus was presented with differently oriented drifting gratings as denoted by the black arrow, with the blue line underneath indicating time of visual stimulus presentation.

NPh_10_4_044304_f006.png

Researchers have adapted voluntary restraint systems for cellular resolution population calcium imaging in behaving rats.34,35,63 These systems used kinematic clamps to achieve high repositioning accuracy and produce the mechanical stability required by cellular resolution imaging [Figs. 6(d)6(f)]. Kinematic clamps131,132 are commonly used in optical and mechanical systems to achieve precise and repeatable alignment. To this end, recent work demonstrates that head fixation devices with micron-scale and submicron-scale repositioning accuracy for cellular resolution imaging are feasible.124 These systems improved upon previously published Kelvin-style kinematic coupling systems63 by utilizing a three vee-groove system, also known as a Maxwell system, which is simpler to manufacture and enables greater long-term performance.133 The design principles described have been scaled up to evaluate voluntary head restraint in larger animals.134

Recent work demonstrates the potential of combining voluntary head-restraint with transgenic rats to record neuron population dynamics over long timescales.34 In this study, a new line of transgenic rats were reported to express GCaMP6f at high levels in hippocampal neurons. These rats were implanted with a newly developed magnetic-based kinematic coupling system and trained in voluntary restraint. Upon becoming proficient, animals performed hundreds of daily fixations over multiple months. 2P imaging through an implanted optical cannula over hippocampal CA1 provided the ability to track a large population of hippocampal neurons for well over a year. Other long term imaging preps (over 140 days) can also be achieved with viral labeling93 (Fig. 7) and with fluorescent dextran (98 days).135 We point out that each of these three groups removed the dura, and future studies will be required to evaluate the impact of different surgical preparations on longitudinal imaging in rats. These studies demonstrate the potential for longitudinal imaging in rats, which could be valuable for experiments on aging, plasticity, and representational drift.

Fig. 7

Longitudinal 2P imaging in rats. (a) Brightfield images of the same cranial window in a rat, beginning from day of implantation (day 0). (b) 2P images of jGCaMP7s-expressing neurons with ROIs and the corresponding spontaneous activity traces from the somatosensory cortex of the same rat in (a). Note that the window quality remained high over 144 days, as reflected in the clarity of the window in the brightfield images.93

NPh_10_4_044304_f007.png

5.

Outlook

Below, we highlight future directions that may improve cellular resolution imaging in the rat model and may help experimentalists determine if the rat model is appropriate for their research program.

5.1.

Next Generation Optical Design

Next generation imaging systems for rats may be improved by increasing the imaging depth, increasing the FOV of imaging systems, and enhancing the SNR to account for the physiological limitations discussed above. The use of 3P imaging can help compensate for the increase in cortical thickness and enable the recording of neuronal activity down to layer 5,38 whereas the use of a large FOV instrument may compensate for the reduction in cell density. The combination of the two, which has been recently described,136,137 could enable activity recording from large neuronal populations in the rat.

Computational approaches have been used to reduce out-of-focus fluorescence neuropil contamination123 and suppress measurement noise in calcium imaging data.138,139 Aside from improving the quality of the data, reduction of noise and out-of-focus light can potentially enable deeper imaging in the rat brain. Computational methods may also aid with the development of new imaging systems. Software designed to simulate the optical, anatomical, and physiological properties of the mouse brain140 may allow for rapid development of next generation imaging systems and provide a standardized ground truth for evaluating their performance. Extending this simulation tool to rats would be a valuable next step and should be feasible given the extensive physiological data available.69,70,80,81

5.2.

Imaging in Cellular Compartments

Several new molecular genetic approaches could be considered in order to improve imaging performance in rats. For example, neuropil contamination could be reduced by expression of soma restricted calcium sensors.141,142 In addition, simultaneous imaging of multiple cell types could be achieved by restricting sensors to readily differentiable cellular compartments, such as axons and soma. Finally, imaging of apical dendrites could allow access to deep cortical neurons, an approach used to support population imaging in macaques.143

5.3.

Multi-Device Imaging

The larger size of the rat loosens spatial constraints with neuroimaging methods. One way would be to incorporate multiple head-mounted microscopes targeting different regions, akin to in vivo electrophysiology. This approach has been applied in mice by targeting two distant regions of interest (ROIs) by developing a smaller one-photon microscope configuration.144,145 Multiple off-the-shelf head-mounted microscopes could be situated on the rat skull using angled, longer relay lenses; this would enable proper clearance for the microscope and lens attachment.

A similar method could be utilized to pair in vivo neuroimaging, in vivo electrophysiology, or perturbation methods in freely behaving rats. As a consequence of a greater working area, ferrules or cannulae could be positioned in areas outside of the imaging window, counter to current methods that record calcium dynamics and provide optogenetic stimulation within the same FOV. Calcium activity of large neural ensembles or neuromodulatory dynamics in one region could be compared with respect to electrophysiological activity—including oscillatory dynamics—in another area.115 Neuroimaging signals in the same FOV could be compared before and after optogenetic or pharmacological manipulations to another structure.

6.

Conclusion

Extending neuroscience tools to a diverse set of species will allow researchers to study how the brains of different species solve similar biological problems.146 This is synergistic with new priorities for cross-species comparative work, in which similar behavioral methods and recording tools are applied across multiple species.147149 Expanding technologies to organisms beyond the species that the technology was originally developed poses a significant challenge. It is our hope that rats can serve both as a valuable model for systems neuroscience and act as a bridge to new framework for applying in vivo imaging tools more broadly across a diverse set of species.

Disclosures

No conflicts of interest, financial or otherwise, are declared by the authors.

Code and Data Availability

Data sharing is not applicable to this article, as no new data were created or analyzed.

Author Contributions

SJK and ASA drafted the manuscript with support from HF, BBS, and ROA. All authors contributed to the editing of the manuscript.

Acknowledgments

SJK is supported by an award from the National Institute on Deafness and Other Communication Disorders (TC32 DC000023). ROA is supported by an award from the National Institutes of Health (NIH Blueprint DSPAN F99NS130925). BBS is supported by NIH award R56MH132732 and a Whitehall Foundation research grant. ASA is supported by NIH award K99 NS119665. We thank Jerry Chen, Jack Giblin, and Aneesh Bal for comments on the manuscript.

References

1. 

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

2. 

M. Kannan et al., “Dual-polarity voltage imaging of the concurrent dynamics of multiple neuron types,” Science, 378 (6619), eabm8797 https://doi.org/10.1126/science.abm8797 SCIEAS 0036-8075 (2022). Google Scholar

3. 

T. Patriarchi et al., “Ultrafast neuronal imaging of dopamine dynamics with designed genetically encoded sensors,” Science, 360 (6396), eaat4422 https://doi.org/10.1126/science.aat4422 SCIEAS 0036-8075 (2018). Google Scholar

4. 

T. Patriarchi et al., “An expanded palette of dopamine sensors for multiplex imaging in vivo,” Nat. Methods, 17 (11), 1147 –1155 https://doi.org/10.1038/s41592-020-0936-3 1548-7091 (2020). Google Scholar

5. 

N. J. Sofroniew et al., “A large field of view two-photon mesoscope with subcellular resolution for in vivo imaging,” eLife, 5 e14472 https://doi.org/10.7554/eLife.14472 (2016). Google Scholar

6. 

C. H. Yu et al., “Diesel2p mesoscope with dual independent scan engines for flexible capture of dynamics in distributed neural circuitry,” Nat. Commun., 12 (1), 6639 https://doi.org/10.1038/s41467-021-26736-4 NCAOBW 2041-1723 (2021). Google Scholar

7. 

W. Zong et al., “Large-scale two-photon calcium imaging in freely moving mice,” Cell, 185 (7), 1240 –1256.e30 https://doi.org/10.1016/j.cell.2022.02.017 CELLB5 0092-8674 (2022). Google Scholar

8. 

J. P. Cunningham and B. M. Yu, “Dimensionality reduction for large-scale neural recordings,” Nat. Neurosci., 17 (11), 1500 –1509 https://doi.org/10.1038/nn.3776 NANEFN 1097-6256 (2014). Google Scholar

9. 

I. H. Stevenson and K. P. Kording, “How advances in neural recording affect data analysis,” Nat. Neurosci., 14 (2), 139 –142 https://doi.org/10.1038/nn.2731 NANEFN 1097-6256 (2011). Google Scholar

10. 

A. E. Urai et al., “Large-scale neural recordings call for new insights to link brain and behavior,” Nat. Neurosci., 25 (1), 11 –19 https://doi.org/10.1038/s41593-021-00980-9 NANEFN 1097-6256 (2022). Google Scholar

11. 

C. Grienberger and A. Konnerth, “Imaging calcium in neurons,” Neuron, 73 (5), 862 –885 https://doi.org/10.1016/j.neuron.2012.02.011 NERNET 0896-6273 (2012). Google Scholar

12. 

G. Grynkiewicz, M. Poenie and R. Y. Tsien, “A new generation of Ca2+ indicators with greatly improved fluorescence properties,” J. Biol. Chem., 260 (6), 3440 –3450 https://doi.org/10.1016/S0021-9258(19)83641-4 JBCHA3 0021-9258 (1985). Google Scholar

13. 

W. Denk, J. H. Strickler and W. W. Webb, “Two-photon laser scanning fluorescence microscopy,” Science, 248 (4951), 73 –76 https://doi.org/10.1126/science.2321027 SCIEAS 0036-8075 (1990). Google Scholar

14. 

A. Miyawaki et al., “Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin,” Nature, 388 (6645), 882 –887 https://doi.org/10.1038/42264 (1997). Google Scholar

15. 

D. A. Dombeck et al., “Imaging large-scale neural activity with cellular resolution in awake, mobile mice,” Neuron, 56 (1), 43 –57 https://doi.org/10.1016/j.neuron.2007.08.003 NERNET 0896-6273 (2007). Google Scholar

16. 

D. A. Dombeck et al., “Functional imaging of hippocampal place cells at cellular resolution during virtual navigation,” Nat. Neurosci., 13 (11), 1433 –1440 https://doi.org/10.1038/nn.2648 NANEFN 1097-6256 (2010). Google Scholar

17. 

K. K. Ghosh et al., “Miniaturized integration of a fluorescence microscope,” Nat. Methods, 8 (10), 871 –878 https://doi.org/10.1038/nmeth.1694 1548-7091 (2011). Google Scholar

18. 

T. W. Chen et al., “Ultrasensitive fluorescent proteins for imaging neuronal activity,” Nature, 499 (7458), 295 –300 https://doi.org/10.1038/nature12354 (2013). Google Scholar

19. 

D. J. Cai et al., “A shared neural ensemble links distinct contextual memories encoded close in time,” Nature, 534 (7605), 115 –118 https://doi.org/10.1038/nature17955 (2016). Google Scholar

20. 

A. Fassihi et al., “Tactile perception and working memory in rats and humans,” Proc. Natl. Acad. Sci. U. S. A., 111 (6), 2331 –2336 https://doi.org/10.1073/pnas.1315171111 (2014). Google Scholar

21. 

A. S. Reinhold et al., “Behavioral and neural correlates of hide-and-seek in rats,” Science, 365 (6458), 1180 –1183 https://doi.org/10.1126/science.aax4705 SCIEAS 0036-8075 (2019). Google Scholar

22. 

R. Poddar, R. Kawai and B. P. Ölveczky, “A fully automated high-throughput training system for rodents,” PLoS One, 8 (12), e83171 https://doi.org/10.1371/journal.pone.0083171 POLNCL 1932-6203 (2013). Google Scholar

23. 

L. Madisen et al., “Transgenic mice for intersectional targeting of neural sensors and effectors with high specificity and performance,” Neuron, 85 (5), 942 –958 https://doi.org/10.1016/j.neuron.2015.02.022 NERNET 0896-6273 (2015). Google Scholar

24. 

N. El-Ayache, J. J. Galligan, “The rat in neuroscience research,” The Laboratory Rat, 1003 –1022 Academic Press( (2020). Google Scholar

25. 

S. A. Barnett, The Story of Rats: Their Impact on Us and Our Impact on Them, Allen and Unwin( (2002). Google Scholar

26. 

J. Altman and G. D. Das, “Autoradiographic and histological evidence of postnatal hippocampal neurogenesis in rats,” J. Comp. Neurol., 124 (3), 319 –335 https://doi.org/10.1002/cne.901240303 JCNEAM 0021-9967 (1965). Google Scholar

27. 

J. O’Keefe and J. Dostrovsky, “The hippocampus as a spatial map: preliminary evidence from unit activity in the freely-moving rat,” Brain Res., 34 171 –175 https://doi.org/10.1016/0006-8993(71)90358-1 BRREAP 0006-8993 (1971). Google Scholar

28. 

J. S. Taube, R. U. Muller and J. B. Ranck, “Head-direction cells recorded from the postsubiculum in freely moving rats. I. Description and quantitative analysis,” J. Neurosci., 10 (2), 420 –435 https://doi.org/10.1523/JNEUROSCI.10-02-00420.1990 JNRSDS 0270-6474 (1990). Google Scholar

29. 

S. Ogawa et al., “Brain magnetic resonance imaging with contrast dependent on blood oxygenation,” Proc. Natl. Acad. Sci. U. S. A., 87 (24), 9868 –9872 https://doi.org/10.1073/pnas.87.24.9868 (1990). Google Scholar

30. 

L. Buck and R. Axel, “A novel multigene family may encode odorant receptors: a molecular basis for odor recognition,” Cell, 65 (1), 175 –187 https://doi.org/10.1016/0092-8674(91)90418-X CELLB5 0092-8674 (1991). Google Scholar

31. 

M. A. Wilson and B. L. McNaughton, “Reactivation of hippocampal ensemble memories during sleep,” Science, 265 (5172), 676 –679 https://doi.org/10.1126/science.8036517 SCIEAS 0036-8075 (1994). Google Scholar

32. 

W. Denk et al., “Anatomical and functional imaging of neurons using 2-photon laser scanning microscopy,” J. Neurosci. Methods, 54 (2), 151 –162 https://doi.org/10.1016/0165-0270(94)90189-9 JNMEDT 0165-0270 (1994). Google Scholar

33. 

T. Hafting et al., “Microstructure of a spatial map in the entorhinal cortex,” Nature, 436 (7052), 801 –806 https://doi.org/10.1038/nature03721 (2005). Google Scholar

34. 

P. D. Rich et al., “Magnetic voluntary head-fixation in transgenic rats enables lifetime imaging of hippocampal neurons,” (2023). Google Scholar

35. 

B. B. Scott et al., “Fronto-parietal cortical circuits encode accumulated evidence with a diversity of timescales,” Neuron, 95 (2), 385 –398.e5 https://doi.org/10.1016/j.neuron.2017.06.013 NERNET 0896-6273 (2017). Google Scholar

36. 

C. Guo et al., “Miniscope-LFOV: a large field of view, single cell resolution, miniature microscope for wired and wire-free imaging of neural dynamics in freely behaving animals,” Sci. Adv., 9 eadg3918 https://doi.org/10.1126/sciadv.adg3918 STAMCV 1468-6996 (2023). Google Scholar

37. 

B. B. Scott et al., “Imaging cortical dynamics in GCaMP transgenic rats with a head-mounted widefield macroscope,” Neuron, 100 (5), 1045 –1058.e5 https://doi.org/10.1016/j.neuron.2018.09.050 NERNET 0896-6273 (2018). Google Scholar

38. 

A. Klioutchnikov et al., “Three-photon head-mounted microscope for imaging deep cortical layers in freely moving rats,” Nat. Methods, 17 (5), 509 –513 https://doi.org/10.1038/s41592-020-0817-9 1548-7091 (2020). Google Scholar

39. 

B. W. Brunton, M. M. Botvinick and C. D. Brody, “Rats and humans can optimally accumulate evidence for decision-making,” Science, 340 (6128), 95 –98 https://doi.org/10.1126/science.1233912 SCIEAS 0036-8075 (2013). Google Scholar

40. 

C. M. Constantinople, A. T. Piet and C. D. Brody, “An analysis of decision under risk in rats,” Curr. Biol., 29 (12), 2066 –2074.e5 https://doi.org/10.1016/j.cub.2019.05.013 CUBLE2 0960-9822 (2019). Google Scholar

41. 

A. Kepecs et al., “Neural correlates, computation and behavioural impact of decision confidence,” Nature, 455 (7210), 227 –231 https://doi.org/10.1038/nature07200 (2008). Google Scholar

42. 

B. B. Scott et al., “Sources of noise during accumulation of evidence in unrestrained and voluntarily head-restrained rats,” eLife, 4 e11308 https://doi.org/10.7554/eLife.11308 (2015). Google Scholar

43. 

N. Uchida, A. Kepecs and Z. F. Mainen, “Seeing at a glance, smelling in a whiff: rapid forms of perceptual decision making,” Nat. Rev. Neurosci., 7 (6), 485 –491 https://doi.org/10.1038/nrn1933 NRNAAN 1471-003X (2006). Google Scholar

44. 

A. Akrami et al., “Posterior parietal cortex represents sensory history and mediates its effects on behavior,” Nature, 554 (7692), 368 –372 https://doi.org/10.1038/nature25510 (2018). Google Scholar

45. 

K. J. Miller, M. M. Botvinick and C. D. Brody, “Dorsal hippocampus contributes to model-based planning,” Nat. Neurosci., 20 (9), 1269 –1276 https://doi.org/10.1038/nn.4613 NANEFN 1097-6256 (2017). Google Scholar

46. 

K. J. Miller, M. M. Botvinick and C. D. Brody, “Value representations in the rodent orbitofrontal cortex drive learning, not choice,” eLife, 11 e64575 https://doi.org/10.7554/eLife.64575 (2022). Google Scholar

47. 

T. S. Critchfield, “Translational contributions of the experimental analysis of behavior,” Behav. Anal., 34 3 –17 https://doi.org/10.1007/BF03392227 (2011). Google Scholar

48. 

J. W. Young and A. Markou, “Translational rodent paradigms to investigate neuromechanisms underlying behaviors relevant to amotivation and altered reward processing in schizophrenia,” Schizophr. Bull., 41 (5), 1024 –1034 https://doi.org/10.1093/schbul/sbv093 SCZBB3 0586-7614 (2015). Google Scholar

49. 

E. G. Patterson-Kane, M. Hunt and D. Harper, “Rats demand social contact,” Anim. Welfare, 11 (3), 327 –332 https://doi.org/10.1017/S0962728600024908 ANWEEF (2002). Google Scholar

50. 

C. Rutte and M. Taborsky, “Generalized reciprocity in rats,” PLoS Biol., 5 (7), e196 https://doi.org/10.1371/journal.pbio.0050196 (2007). Google Scholar

51. 

D. S. Viana et al., “Cognitive and motivational requirements for the emergence of cooperation in a rat social game,” PLoS One, 5 (1), e8483 https://doi.org/10.1371/journal.pone.0008483 POLNCL 1932-6203 (2010). Google Scholar

52. 

I. B. A. Bartal, J. Decety and P. Mason, “Empathy and pro-social behavior in rats,” Science, 334 (6061), 1427 –1430 https://doi.org/10.1126/science.1210789 SCIEAS 0036-8075 (2011). Google Scholar

53. 

M. Nagy et al., “Synergistic benefits of group search in rats,” Curr. Biol., 30 4733 –4738.e4 https://doi.org/10.1016/j.cub.2020.08.079 CUBLE2 0960-9822 (2020). Google Scholar

54. 

V. Pedrosa et al., “Humans, rats and mice show species-specific adaptations to sensory statistics in categorisation behavior,” (2023). Google Scholar

55. 

I. Q. Whishaw et al., “Accelerated nervous system development contributes to behavioral efficiency in the laboratory mouse: a behavioral review and theoretical proposal,” Dev. Psychobiol., 39 (3), 151 –170 https://doi.org/10.1002/dev.1041 DEPBA5 0012-1630 (2001). Google Scholar

56. 

V. Vengeliene, A. Bilbao and R. Spanagel, “The alcohol deprivation effect model for studying relapse behavior: a comparison between rats and mice,” Alcohol, 48 (3), 313 –320 https://doi.org/10.1016/j.alcohol.2014.03.002 ALCOEX 0741-8329 (2014). Google Scholar

57. 

D. Aronov and D. W. Tank, “Engagement of neural circuits underlying 2D spatial navigation in a rodent virtual reality system,” Neuron, 84 (2), 442 –456 https://doi.org/10.1016/j.neuron.2014.08.042 NERNET 0896-6273 (2014). Google Scholar

58. 

C. Holscher et al., “Rats are able to navigate in virtual environments,” J. Exp. Biol., 208 (3), 561 –569 https://doi.org/10.1242/jeb.01371 JEBIAM 0022-0949 (2005). Google Scholar

59. 

K. Safaryan and M. R. Mehta, “Enhanced hippocampal theta rhythmicity and emergence of eta oscillation in virtual reality,” Nat. Neurosci., 24 (8), 1065 –1070 https://doi.org/10.1038/s41593-021-00871-z NANEFN 1097-6256 (2021). Google Scholar

60. 

T. J. Bussey et al., “The touchscreen cognitive testing method for rodents: how to get the best out of your rat,” Learn. Mem., 15 (7), 516 –523 https://doi.org/10.1101/lm.987808 (2008). Google Scholar

61. 

S. V. Girman, “Means of restricting the movements of conscious rats in neurophysiologic experiments,” Zh. Vyssh. Nerv. Deiat. Im. IP Pavlova, 30 (5), 1087 –1089 (1980). Google Scholar

62. 

S. V. Girman, “Responses of neurons of primary visual cortex of awake unrestrained rats to visual stimuli,” Neurosci. Behav. Physiol., 15 379 –386 https://doi.org/10.1007/BF01184022 NBHPBT 0097-0549 (1985). Google Scholar

63. 

B. B. Scott, C. D. Brody and D. W. Tank, “Cellular resolution functional imaging in behaving rats using voluntary head restraint,” Neuron, 80 (2), 371 –384 https://doi.org/10.1016/j.neuron.2013.08.002 NERNET 0896-6273 (2013). Google Scholar

64. 

K. Luxem et al., “Open-source tools for behavioral video analysis: setup, methods, and best practices,” eLife, 12 e79305 https://doi.org/10.7554/eLife.79305 (2023). Google Scholar

65. 

Charles River Laboratories, Long Evans Rat, https://www.criver.com/products-services/find-model/long-evans-rat (19 November 2023). Google Scholar

66. 

J. Sawinski et al., “Visually evoked activity in cortical cells imaged in freely moving animals,” Proc. Natl. Acad. Sci. U. S. A., 106 (46), 19557 –19562 https://doi.org/10.1073/pnas.0903680106 (2009). Google Scholar

67. 

J. Y. Rhee et al., “Neural correlates of visual object recognition in rats,” (2023). Google Scholar

68. 

C. Schwarz et al., “The head-fixed behaving rat—procedures and pitfalls,” Somatosens. Mot. Res., 27 (4), 131 –148 https://doi.org/10.3109/08990220.2010.513111 SMOREZ 1369-1651 (2010). Google Scholar

69. 

S. Herculano-Houzel, B. Mota and R. Lent, “Cellular scaling rules for rodent brains,” Proc. Natl. Acad. Sci. U. S. A., 103 (32), 12138 –12143 https://doi.org/10.1073/pnas.0604911103 (2006). Google Scholar

70. 

S. Herculano-Houzel et al., “Mammalian brains are made of these: a dataset of the numbers and densities of neuronal and nonneuronal cells in the brain of glires, primates, scandentia, eulipotyphlans, afrotherians and artiodactyls, and their relationship with body mass,” Brain Behav. Evol., 86 (3–4), 145 –163 https://doi.org/10.1159/000437413 (2015). Google Scholar

71. 

A. J. Rockel, R. W. Hiorns and T. P. Powell, “The basic uniformity in structure of the neocortex,” Brain, 103 (2), 221 –244 https://doi.org/10.1093/brain/103.2.221 (1980). Google Scholar

72. 

C. Beaulieu, “Numerical data on neocortical neurons in adult rat, with special reference to the GABA population,” Brain Res., 609 (1–2), 284 –292 https://doi.org/10.1016/0006-8993(93)90884-P BRREAP 0006-8993 (1993). Google Scholar

73. 

A. Morales-Gregorio, A. van Meegen and S. J. van Albada, “Ubiquitous lognormal distribution of neuron densities in mammalian cerebral cortex,” Cerebral Cortex, 33 (16), 9439 –9449 https://doi.org/10.1093/cercor/bhad160 (2023). Google Scholar

74. 

M. Azimipour et al., “Extraction of optical properties and prediction of light distribution in rat brain tissue,” J. Biomed. Opt., 19 (7), 075001 https://doi.org/10.1117/1.JBO.19.7.075001 JBOPFO 1083-3668 (2014). Google Scholar

75. 

D. Kleinfeld and W. Denk, Two-Photon Imaging of Cortical Microcirculation. Imaging in Neuroscience and Development: A Laboratory Manual, 701 –705 Cold Spring Harbor Laboratory Press, Cold Spring Harbor (2000). Google Scholar

76. 

M. Oheim et al., “Two-photon microscopy in brain tissue: parameters influencing the imaging depth,” J. Neurosci. Methods, 111 (1), 29 –37 https://doi.org/10.1016/S0165-0270(01)00438-1 JNMEDT 0165-0270 (2001). Google Scholar

77. 

D. Kobat et al., “Deep tissue multiphoton microscopy using longer wavelength excitation,” Opt. Express, 17 (16), 13354 –13364 https://doi.org/10.1364/OE.17.013354 OPEXFF 1094-4087 (2009). Google Scholar

78. 

T. Wang et al., “Quantitative analysis of 1300-nm three-photon calcium imaging in the mouse brain,” eLife, 9 e53205 https://doi.org/10.7554/eLife.53205 (2020). Google Scholar

79. 

K. Svoboda et al., “Spread of dendritic excitation in layer 2/3 pyramidal neurons in rat barrel cortex in vivo,” Nat. Neurosci., 2 (1), 65 –73 https://doi.org/10.1038/4569 NANEFN 1097-6256 (1999). Google Scholar

80. 

P. Blinder et al., “The cortical angiome: an interconnected vascular network with noncolumnar patterns of blood flow,” Nat. Neurosci., 16 (7), 889 –897 https://doi.org/10.1038/nn.3426 NANEFN 1097-6256 (2013). Google Scholar

81. 

P.S. Tsai et al., “Correlations of neuronal and microvascular densities in murine cortex revealed by direct counting and colocalization of nuclei and vessels,” J. Neurosci., 29 (46), 14553 –14570 https://doi.org/10.1523/JNEUROSCI.3287-09.2009 JNRSDS 0270-6474 (2009). Google Scholar

82. 

V. V. Barun and A. P. Ivanov, “Estimate of the contribution of localized light absorption by blood vessels to the optical properties of biological tissue,” Opt. Spectrosc., 96 940 –945 https://doi.org/10.1134/1.1771432 OPSUA3 0030-400X (2004). Google Scholar

83. 

E. Chaigneau et al., “Two-photon imaging of capillary blood flow in olfactory bulb glomeruli,” Proc. Natl. Acad. Sci. U. S. A., 100 13081 –13086 https://doi.org/10.1073/pnas.2133652100 (2003). Google Scholar

84. 

D. Wang et al., “Biocompatible and photostable AIE dots with red emission for in vivo two-photon bioimaging,” Sci. Rep., 4 (1), 4279 https://doi.org/10.1038/srep04279 (2014). Google Scholar

85. 

M. Vedi et al., “2022 updates to the rat genome database: a findable, accessible, interoperable, and reuslable (FAIR) resource,” Genetics, 224 iyad042 https://doi.org/10.1093/genetics/iyad042 GENTAE 0016-6731 (2023). Google Scholar

86. 

J. R. Pettibone et al., “Knock-in rat lines with cre recombinase at the dopamine D1 and adenosine 2a receptor loci,” eNeuro, 6 (5), ENEURO.0163-19.2019 https://doi.org/10.1523/ENEURO.0163-19.2019 (2019). Google Scholar

87. 

G. D. Stuber, A. M. Stamatakis and P. A. Kantak, “Considerations when using cre-driver rodent lines for studying ventral tegmental area circuitry,” Neuron, 85 (2), 439 –445 https://doi.org/10.1016/j.neuron.2014.12.034 NERNET 0896-6273 (2015). Google Scholar

88. 

I. B. Witten et al., “Recombinase-driver rat lines: tools, techniques, and optogenetic application to dopamine-mediated reinforcement,” Neuron, 72 (5), 721 –733 https://doi.org/10.1016/j.neuron.2011.10.028 NERNET 0896-6273 (2011). Google Scholar

89. 

, “Simons Foundation Autism Research Initiative (SFARI) rat models,” https://www.sfari.org/resource/rat-models/ (2022). Google Scholar

90. 

Rat Genome Database, https://rgd.mcw.edu/wg/resource-links/#rat_resources (19 November 2023). Google Scholar

91. 

J. M. Gee et al., “Imaging activity in astrocytes and neurons with genetically encoded calcium indicators following in utero electroporation,” Front. Mol. Neurosci., 8 10 https://doi.org/10.3389/fnmol.2015.00010 (2015). Google Scholar

92. 

A. S. Alexander, B. B. Scott and M. E. Hasselmo, “Laminar and projection-specific calcium imaging of spatial memory related retrosplenial cortex dynamics in freely moving rats,” in Prog. No 866.05. Neurosci. Meeting Planner, (2021). Google Scholar

93. 

S. Chornyy et al., “Cellular-resolution monitoring of ischemic stroke pathologies in the rat cortex,” Biomed. Opt. Express, 12 (8), 4901 –4919 https://doi.org/10.1364/BOE.432688 BOEICL 2156-7085 (2021). Google Scholar

94. 

C. M. Cameron et al., “Increased cocaine motivation is associated with degraded spatial and temporal representations in IL-NAc neurons,” Neuron, 103 (1), 80 –91.e7 https://doi.org/10.1016/j.neuron.2019.04.015 NERNET 0896-6273 (2019). Google Scholar

95. 

F. Gobbo et al., “Neuronal signature of spatial decision-making during navigation by freely moving rats by using calcium imaging,” Proc. Natl. Acad. Sci. U. S. A., 119 (44), e2212152119 https://doi.org/10.1073/pnas.2212152119 (2022). Google Scholar

96. 

E. E. Hart et al., “Chemogenetic modulation and single-photon calcium imaging in anterior cingulate cortex reveal a mechanism for effort-based decisions,” J. Neurosci., 40 (29), 5628 –5643 https://doi.org/10.1523/JNEUROSCI.2548-19.2020 JNRSDS 0270-6474 (2020). Google Scholar

97. 

H. S. Wirtshafter and J. F. Disterhoft, “In vivo multi-day calcium imaging of CA1 hippocampus in freely moving rats reveals a high preponderance of place cells with consistent place fields,” J. Neurosci., 42 (22), 4538 –4554 https://doi.org/10.1523/JNEUROSCI.1750-21.2022 JNRSDS 0270-6474 (2022). Google Scholar

98. 

H. S. Wirtshafter and J. F. Disterhoft, “Place cells are nonrandomly clustered by field location in CA1 hippocampus,” Hippocampus, 33 (2), 65 –84 https://doi.org/10.1002/hipo.23489 (2023). Google Scholar

99. 

T. L. Daigle et al., “A suite of transgenic driver and reporter mouse lines with enhanced brain-cell-type targeting and functionality,” Cell, 174 (2), 465 –480.e22 https://doi.org/10.1016/j.cell.2018.06.035 CELLB5 0092-8674 (2018). Google Scholar

100. 

H. K. Decot et al., “Coordination of brain-wide activity dynamics by dopaminergic neurons,” Neuropsychopharmacology, 42 (3), 615 –627 https://doi.org/10.1038/npp.2016.151 NEROEW 0893-133X (2017). Google Scholar

101. 

R. C. Challis et al., “Systemic AAV vectors for widespread and targeted gene delivery in rodents,” Nat. Protoc., 14 (2), 379 –414 https://doi.org/10.1038/s41596-018-0097-3 1754-2189 (2019). Google Scholar

102. 

R. D. Dayton, M. S. Grames and R. L. Klein, “More expansive gene transfer to the rat CNS: AAV PHP. EB vector dose–response and comparison to AAV PHP.B,” Gene Ther., 25 (5), 392 –400 https://doi.org/10.1038/s41434-018-0028-5 GETHEC 0969-7128 (2018). Google Scholar

103. 

K. L. Jackson, R. D. Dayton and R. L. Klein, “AAV9 supports wide-scale transduction of the CNS and TDP-43 disease modeling in adult rats,” Mol. Ther. Methods Clin. Dev., 2 15036 https://doi.org/10.1038/mtm.2015.36 (2015). Google Scholar

104. 

D. Chatterjee et al., “Enhanced CNS transduction from AAV. PHP. eB infusion into the cisterna magna of older adult rats compared to AAV9,” Gene Ther., 29 (6), 390 –397 https://doi.org/10.1038/s41434-021-00244-y GETHEC 0969-7128 (2022). Google Scholar

105. 

K. L. Pietersz et al., “Transduction patterns in the CNS following various routes of AAV-5-mediated gene delivery,” Gene Ther., 28 (7–8), 435 –446 https://doi.org/10.1038/s41434-020-0178-0 GETHEC 0969-7128 (2021). Google Scholar

106. 

K. L. Jackson et al., “Better targeting, better efficiency for wide-scale neuronal transduction with the synapsin promoter and AAV-PHP.B,” Front. Mol. Neurosci., 9 116 https://doi.org/10.3389/fnmol.2016.00116 (2016). Google Scholar

107. 

D. J. Garrett et al., “In utero recombinant adeno-associated virus gene transfer in mice, rats, and primates,” BMC Biotech., 3 16 https://doi.org/10.1186/1472-6750-3-16 (2003). Google Scholar

108. 

T. Saito and N. Nakatsuji, “Efficient gene transfer into the embryonic mouse brain using in vivo electroporation,” Dev. Biol., 240 (1), 237 –246 https://doi.org/10.1006/dbio.2001.0439 DEBIAO 0012-1606 (2001). Google Scholar

109. 

J. Szczurkowska et al., “Targeted in vivo genetic manipulation of the mouse or rat brain by in utero electroporation with a triple-electrode probe,” Nat. Protoc., 11 399 –412 https://doi.org/10.1038/nprot.2016.014 1754-2189 (2016). Google Scholar

110. 

H. Tabata and K. Nakajima, “Efficient in utero gene transfer system to the developing mouse brain using electroporation: visualization of neuronal migration in the developing cortex,” Neuroscience, 103 (4), 865 –872 https://doi.org/10.1016/S0306-4522(01)00016-1 (2001). Google Scholar

111. 

W. Walantus, L. Elias and A. Kriegstein, “In utero intraventricular injection and electroporation of E16 rat embryos,” J. Vis. Exp., (6), e236 https://doi.org/10.3791/236 (2007). Google Scholar

112. 

L. Chansel-Debordeaux et al., “In utero delivery of rAAV2/9 induces neuronal expression of the transgene in the brain: towards new models of Parkinson’s disease,” Gene Ther., 24 (12), 801 –809 https://doi.org/10.1038/gt.2017.84 GETHEC 0969-7128 (2017). Google Scholar

113. 

J. M. Rosin and D. M. Kurrasch, “In utero electroporation induces cell death and alters embryonic microglia morphology and expression signatures in the developing hypothalamus,” J. Neuroinflammation, 15 (1), 181 https://doi.org/10.1186/s12974-018-1213-6 (2018). Google Scholar

114. 

D. Aharoni and T. M. Hoogland, “Circuit investigations with open-source miniaturized microscopes: past, present and future,” Front. Cell Neurosci., 13 141 https://doi.org/10.3389/fncel.2019.00141 (2019). Google Scholar

115. 

N. R. Kinsky, J. Haddad and K. Diba, “Combined electrophysiology and imaging to investigate hippocampal-cortical interactions during memory consolidation,” in Soc. Neurosci. Abstr., (2022). Google Scholar

116. 

K. Saxena et al., “iHELMET: a 3D-printing solution for safe endoscopic Ca2+ recording in social neuroscience,” J. Neurosci. Methods, 355 109109 https://doi.org/10.1016/j.jneumeth.2021.109109 JNMEDT 0165-0270 (2021). Google Scholar

117. 

M. L. Andermann et al., “Chronic cellular imaging of entire cortical columns in awake mice using microprisms,” Neuron, 80 (4), 900 –913 https://doi.org/10.1016/j.neuron.2013.07.052 NERNET 0896-6273 (2013). Google Scholar

118. 

S. Gulati, V. Y. Cao and S. Otte, “Multi-layer cortical Ca2+ imaging in freely moving mice with prism probes and miniaturized fluorescence microscopy,” J. Vis. Exp., 124 e55579 https://doi.org/10.3791/55579 (2017). Google Scholar

119. 

Z. Chen et al., “A hardware system for real-time decoding of in vivo calcium imaging data,” eLife, 12 e78344 https://doi.org/10.7554/eLife.78344 (2023). Google Scholar

120. 

F. Helmchen et al., “A miniature head-mounted two-photon microscope: high-resolution brain imaging in freely moving animals,” Neuron, 31 (6), 903 –912 https://doi.org/10.1016/S0896-6273(01)00421-4 NERNET 0896-6273 (2001). Google Scholar

121. 

A. Klioutchnikov et al., “A three-photon head-mounted microscope for imaging all layers of visual cortex in freely moving mice,” Nat. Methods, 20 610 –616 https://doi.org/10.1038/s41592-022-01688-9 1548-7091 (2022). Google Scholar

122. 

F. Helmchen and W. Denk, “Deep tissue two-photon microscopy,” Nat. Methods, 2 (12), 932 –940 https://doi.org/10.1038/nmeth818 1548-7091 (2005). Google Scholar

123. 

A. Glas et al., “Benchmarking miniaturized microscopy against two-photon calcium imaging using single-cell orientation tuning in mouse visual cortex,” PLoS One, 14 (4), e0214954 https://doi.org/10.1371/journal.pone.0214954 POLNCL 1932-6203 (2019). Google Scholar

124. 

S. J. Kim, A. H. Slocum and B. B. Scott, “A miniature kinematic coupling device for mouse head fixation,” J. Neurosci. Methods, 372 109549 https://doi.org/10.1016/j.jneumeth.2022.109549 JNMEDT 0165-0270 (2022). Google Scholar

125. 

M. L. Culpepper, “Design of quasi-kinematic couplings,” Precis. Eng., 28 (3), 338 –357 https://doi.org/10.1016/j.precisioneng.2002.12.001 PREGDL 0141-6359 (2004). Google Scholar

126. 

A. R. Kampff et al., “The voluntary head-restrained rat,” in Program No. 819.18. 2010 Neurosci. Meeting Planner, (2010). Google Scholar

127. 

R. Aoki et al., “An automated platform for high-throughput mouse behavior and physiology with voluntary head-fixation,” Nat. Commun., 8 (1), 1196 https://doi.org/10.1038/s41467-017-01371-0 NCAOBW 2041-1723 (2017). Google Scholar

128. 

Y. Hao, A. M. Thomas and N. Li, “Fully autonomous mouse behavioral and optogenetic experiments in home-cage,” eLife, 10 e66112 https://doi.org/10.7554/eLife.66112 (2021). Google Scholar

129. 

T. H. Murphy et al., “High-throughput automated home-cage mesoscopic functional imaging of mouse cortex,” Nat. Commun., 7 (1), 11611 https://doi.org/10.1038/ncomms11611 NCAOBW 2041-1723 (2016). Google Scholar

130. 

T. H. Murphy et al., “Automated task training and longitudinal monitoring of mouse mesoscale cortical circuits using home cages,” eLife, 9 e55964 https://doi.org/10.7554/eLife.55964 (2020). Google Scholar

131. 

C. Evans, Precision Engineering: An Evolutionary View, Cranfield Press, Bedford (1989). Google Scholar

132. 

J. C. Maxwell and W. D. Niven, “General considerations concerning scientific apparatus,” Sci. Pap. JC Maxwell, 2 507 –508 (1890). Google Scholar

133. 

A. H. Slocum, “Design of three-groove kinematic couplings,” Precis. Eng., 14 (2), 67 –76 https://doi.org/10.1016/0141-6359(92)90051-W PREGDL 0141-6359 (1992). Google Scholar

134. 

J. D. Walker et al., “A platform for semiautomated voluntary training of common marmosets for behavioral neuroscience,” J. Neurophysiol., 123 (4), 1420 –1426 https://doi.org/10.1152/jn.00300.2019 JONEA4 0022-3077 (2020). Google Scholar

135. 

M. M. Koletar et al., “Refinement of a chronic cranial window implant in the rat for longitudinal in vivo two-photon fluorescence microscopy of neurovascular function,” Sci. Rep., 9 (1), 5499 https://doi.org/10.1038/s41598-019-41966-9 (2019). Google Scholar

136. 

A. T. Mok et al., “A large field of view two-and three-photon microscope for high-resolution deep tissue imaging,” in CLEO: Appl. and Technol., ATh5A-1 (2023). Google Scholar

137. 

N. R. Shvedov et al., “Deep brain three-photon imaging in transgenic songbirds,” in Program No. PSTR090.14. 2023 Neurosci. Meeting Planner, (2023). Google Scholar

138. 

X. Li et al., “Reinforcing neuron extraction and spike inference in calcium imaging using deep self-supervised denoising,” Nat. Methods, 18 (11), 1395 –1400 https://doi.org/10.1038/s41592-021-01225-0 1548-7091 (2021). Google Scholar

139. 

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

140. 

A. Song et al., “Neural anatomy and optical microscopy (NAOMi) simulation for evaluating calcium imaging methods,” J. Neurosci. Methods, 358 109173 https://doi.org/10.1016/j.jneumeth.2021.109173 JNMEDT 0165-0270 (2021). Google Scholar

141. 

Y. Chen et al., “Soma-targeted imaging of neural circuits by ribosome tethering,” Neuron, 107 (3), 454 –469.e6 https://doi.org/10.1016/j.neuron.2020.05.005 NERNET 0896-6273 (2020). Google Scholar

142. 

S. Grødem et al., “An updated suite of viral vectors for in vivo calcium imaging using intracerebral and retro-orbital injections in male mice,” Nat. Commun., 14 (1), 608 https://doi.org/10.1038/s41467-023-36324-3 NCAOBW 2041-1723 (2023). Google Scholar

143. 

E. M. Trautmann et al., “Dendritic calcium signals in rhesus macaque motor cortex drive an optical brain-computer interface,” Nat. Commun., 12 (1), 3689 https://doi.org/10.1038/s41467-021-23884-5 NCAOBW 2041-1723 (2021). Google Scholar

144. 

A. de Groot et al., “NINscope, a versatile miniscope for multi-region circuit investigations,” eLife, 9 e49987 https://doi.org/10.7554/eLife.49987 (2020). Google Scholar

145. 

W. G. Gonzalez et al., “Persistence of neuronal representations through time and damage in the hippocampus,” Science, 365 (6455), 821 –825 https://doi.org/10.1126/science.aav9199 SCIEAS 0036-8075 (2019). Google Scholar

146. 

G. Laurent, “On the value of model diversity in neuroscience,” Nat. Rev. Neurosci., 21 (8), 395 –396 https://doi.org/10.1038/s41583-020-0323-1 NRNAAN 1471-003X (2020). Google Scholar

147. 

D. Badre, M. J. Frank and C. I. Moore, “Interactionist neuroscience,” Neuron, 88 855 –860 https://doi.org/10.1016/j.neuron.2015.10.021 NERNET 0896-6273 (2015). Google Scholar

148. 

H. C. Barron et al., “Cross-species neuroscience: closing the explanatory gap,” Philos. Trans. R. Soc. B Biol. Sci., 376 20190633 https://doi.org/10.1098/rstb.2019.0633 (2021). Google Scholar

149. 

M. M. Nour, Y. Liu and R. J. Dolan, “Functional neuroimaging in psychiatry and the case for failing better,” Neuron, 110 (16), 2524 –2544 https://doi.org/10.1016/j.neuron.2022.07.005 NERNET 0896-6273 (2022). Google Scholar

Biographies of the authors are not available.

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.
Su Jin Kim, Rifqi O. Affan, Hadas Frostig, Benjamin B. Scott, and Andrew S. Alexander "Advances in cellular resolution microscopy for brain imaging in rats," Neurophotonics 10(4), 044304 (30 November 2023). https://doi.org/10.1117/1.NPh.10.4.044304
Received: 5 May 2023; Accepted: 7 November 2023; Published: 30 November 2023
Advertisement
Advertisement
KEYWORDS
Microscopes

Imaging systems

Calcium

Head

Neurons

Image resolution

Biological imaging

Back to Top