Open Access
10 September 2022 Multiscale photoacoustic tomography of neural activities with GCaMP calcium indicators
Ruiying Zhang, Lei S. Li, Bin Rao, Haoyang Rong, Min-Yu Sun, Junjie Yao, Ruimin Chen, Qifa Zhou, Steven Mennerick, Baranidharan Raman, Lihong V. Wang
Author Affiliations +
Abstract

Significance: Optical imaging of responses in fluorescently labeled neurons has progressed significantly in recent years. However, there is still a need to monitor neural activities at divergent spatial scales and at depths beyond the optical diffusion limit.

Aim: To meet these needs, we aim to develop multiscale photoacoustic tomography (PAT) to image neural activities across spatial scales with a genetically encoded calcium indicator GCaMP.

Approach: First, using photoacoustic microscopy, we show that depth-resolved GCaMP signals can be monitored in vivo from a fly brain in response to odor stimulation without depth scanning and even with the cuticle intact. In vivo monitoring of GCaMP signals was also demonstrated in mouse brains. Next, using photoacoustic computed tomography, we imaged neural responses of a mouse brain slice at depths beyond the optical diffusion limit.

Results: We provide the first unambiguous demonstration that multiscale PAT can be used to record neural activities in transgenic flies and mice with select neurons expressing GCaMP.

Conclusions: Our results indicate that the combination of multiscale PAT and fluorescent neural activity indicators provides a methodology for imaging targeted neurons at various scales.

1.

Introduction

Monitoring the neural activity of large populations of neurons with fine spatial and temporal resolution is arguably one of the most important challenges in neuroscience.13 To achieve this goal, several multiunit electrophysiological recording techniques46 and optical imaging methodologies79 have been developed in recent decades. Electrophysiological recording methods are still considered the gold standard because they directly monitor intracellular or extracellular neural signals from individual neurons or small groups of neurons with high sensitivity and temporal fidelity.10 Complementing these techniques, noninvasive optical imaging techniques that allow for larger spatial coverage with finer spatial resolution have been used to study computations performed by neural networks.7,11,12 Combined with fluorescent calcium and voltage indicators, multiphoton microscopy, including wide-field two-photon microscopy,1315 Bessel-beam two-photon microscopy1618 and three-photon microscopy,1922 has enabled optical mapping of neural circuits from hundreds to thousands of neurons, with high spatiotemporal resolution in optically scattering brain tissue.23,24 However, several limitations still exist. For example, the point-by-point scanning of confocal microscopy and two-photon microscopy creates a trade-off between spatial coverage and temporal resolution. More importantly, light scattering in the neural tissue fundamentally restricts the current optical neural imaging methods to the superficial tissue layer (1  mm).25,26 To address this issue, invasive techniques using deeply implanted fiber-optic probes have been recently used to image nonsuperficial regions inside mouse brains.2730 Therefore, there is a strong need for noninvasive optical imaging methods that can break the optical diffusion limit and provide rapid depth-resolved monitoring of neural activity.

Alternative to conventional optical imaging methods, photoacoustic tomography (PAT) has been demonstrated as a powerful brain imaging technique capable of providing high-resolution images with optical absorption contrast at depths far beyond the optical diffusion limit.3134 Briefly, PAT is based on the photoacoustic (PA) effect, which starts with optical absorption by molecules in tissue and ends with ultrasonic emission through thermoelastic expansion.35 Similar to ultrasound imaging, PAT has inherent depth-resolved imaging capability based on the acoustic arrival time or acoustic focusing. Because PAT is sensitive to optical absorption contrast (both endogenous and exogenous), it is well suited for functional, molecular, and metabolic brain imaging.3640 In addition, PAT is highly scalable in its spatial resolution, penetration depth and imaging speed. Previous endeavors in PA brain studies have focused on measuring hemodynamic parameters, including the blood vessel diameter, blood flow, total hemoglobin concentration, and oxygenation saturation of hemoglobin,4150 which are usually slow and indirect indicators of neural activities through neurovascular coupling.51 In contrast, intracellular calcium signals respond to neural activities faster than hemodynamic signals and can be recorded optically with the help of intracellular calcium indicators.52 Among all of the calcium indicators, GCaMP, with its relatively fast conformational change upon calcium binding during action potentials, is by far the most widely used genetically encoded calcium-sensitive protein.53,54 The conformational change of GCaMP results in a substantial increase in its optical absorption, which is ideal for PAT imaging of calcium signals with high sensitivity. However, it is yet to be determined whether experimental results support this hypothesis and how the PA signals compare to the fluorescence signals.

Here, for the first time, we clearly demonstrate that PAT can be used to monitor neural activities in select neurons expressing GCaMP indicators in transgenic invertebrate (fly) and vertebrate (mouse) models. To accommodate different spatial scales, we adapted two PAT embodiments—photoacoustic microscopy (PAM) and photoacoustic computed tomography (PACT). With PAM, we demonstrate high-resolution depth-resolved neural imaging without depth scanning, which is desired for monitoring neural activities with high temporal resolution. With PACT, we demonstrate wide-field imaging of GCaMP calcium signals at depths beyond the optical diffusion limit. The combination of PAT and GCaMP or other potential calcium indicators provides a promising platform for functional brain imaging at different spatial scales.

2.

Materials and Methods

2.1.

PAT of Neuronal Calcium Signals with GCaMP

In PAT, as tissues are excited by laser pulses, laser energy is absorbed by biomolecules, and converted through both the fluorescence channel by radiative relaxation and the PA channel by nonradiative relaxation. An ultrasonic transducer or transducer array is used to detect the PA waves as they propagate out of the tissue. The detected PA signals are used to reconstruct an image that maps the original optical energy deposition due to nonradiative relaxation in the tissue.

Based on the image formation methods, PAT has two major implementations. The first combines mechanical scanning of a focused single-element ultrasonic transducer with direct image formation and is commonly used in PAM. The second combines parallel detection by a multielement ultrasonic transducer array or a mechanical/electronic scanning equivalent with image reconstruction and is used in PACT. In both PAM and PACT, the PA amplitude is proportional to the optical absorption coefficient of the absorber in the unit of cm1, its nonradiative quantum yield (approximately equal to 1 – the fluorescence quantum yield if the photochemical channel is negligible), and the local optical fluence (or exposure) in the unit of J/m2.

Two GCaMP variants were used in this study: GCaMP5G for fly brain imaging and GCaMP6f for mouse brain slice imaging. In general, GCaMP consists of a circularly permuted EGFP, an M13 fragment of myosin light chain kinase, and a calcium binding site of calmodulin. Action potentials trigger a large and rapid influx of Ca2+.55 In the presence of calcium, calmodulin interacts with the M13 fragment, eliciting a conformational change in EGFP and resulting in an increase in its optical absorption coefficient.56,57 Because the PA signal of GCaMP is proportional to its absorption coefficient, PAT has a 100% sensitivity to the absorption change of GCaMP upon the binding of Ca2+.

2.2.

PAM of the Fly Brain in Vivo

A dual-modality PA and epifluorescence microscopy system shown schematically in Fig. 1(a) was used to image the fly brain. A tunable OPO laser (a 1-kHz pulse repetition rate, 5-ns pulse width; NT242-SH, Ekspla, Ltd.) provided 488-nm wavelength excitation light, which was collimated and coupled into a miniature imaging probe containing an achromatic focusing lens with a 13-mm focal length and a customized ring-shaped ultrasonic transducer (a 40-MHz central frequency, 100% bandwidth, and 6.2-mm focal length). The laser beam formed a 6-μm optical focal spot confocally aligned with the acoustic focus of the ring transducer. The axial resolution of the PAM, determined by the ultrasonic transducer bandwidth, was calculated to be 21  μm.58 A beam splitter reflected the fluorescent light for fluorescence imaging. The fluorescent light passed through another focusing lens (a 50-mm focal length) and a bandpass filter (a 525-nm central wavelength, 39-nm bandwidth; Thorlabs, Inc.) before being collected by an EMCCD camera (iXon Ultra 888, Andor, Ltd.). PA signals were sampled by a 200-MHz data acquisition (DAQ) system (National Instruments, Corp.), and fluorescence signals were recorded by the EMCCD camera’s onboard DAQ system. For each M-mode recording, PA signals before and after odor stimulation from a volume (6  μm×6  μm×200  μm) of neurons inside the brain tissue were acquired at a one-dimensional (1D) imaging rate of 1 kHz. To form a three-dimensional (3D) image of the fly brain, two-dimensional (2D) raster scanning was performed by motor stages. A photodiode (not shown) was used to compensate for the laser pulse energy fluctuation.

Fig. 1

PAM of the GCaMP5G-expressing fly brain. (a) Schematic of the dual-modality PAM and epifluorescence microscopy (EFM) system. PAM and EFM share the same light path for excitation. The photoacoustic (PA) signal is detected by a ring-shaped ultrasonic transducer, whereas the fluorescence signal is detected by an EMCCD. (b) PAM image of a fly with endogenous contrast, showing the major parts of the fly, including the antennae, compound eyes, head, and thorax. The PA signals mostly come from pigments (e.g., melanin) on the fly cuticle.

JBO_27_9_096004_f001.png

2.3.

Fly Preparation and Odor Stimulation

GCaMP5G flies were purchased from the Bloomington Drosophila Stock Center at Indiana University. Flies were anesthetized by CO2 until their movement stopped and were moved to a customized fly mount (a block setup59 or a foil setup60). For the block setup, the fly was mounted and head-fixed using two-component silicone glue (KWIK-SIL, World Precision Instruments, Inc.). A thin wire was used to pull the antennae plate forward and make space between the antennae plate and the head. A small rectangular hole was created on the bottom of a 60  mm×15  mm petri dish (Falcon®, Corning, Inc.). Four sections of double-sided tape were used to create a rectangular window slightly larger than the head of the fly. The petri dish with the small rectangular open window at the bottom was placed on top of the fly’s head and fixed onto the mounting block by double-sided tape. The interior rectangular edges were then sealed with two-component silicone glue. For the foil setup, a piece of aluminum foil was glued to the top of a supporting plate. A 60  mm×15  mm Petri dish (Falcon®, Corning, Inc.) with its bottom removed was attached to the top of the supporting plate to form a saline chamber. A hole was made in the foil according to the fly’s size. The fly was then wedged into the hole with its dorsal side facing up. The gap between the fly and foil was sealed using epoxy (Permatex, Inc.). The ventral side, especially the antennae, of the fly was kept underneath the foil. The fly was then transferred to the PA/fluorescence imaging setup for experiments with the cuticle intact.

For experiments with the cuticle removed, a fine forceps (Dumont #5sf, Fine Science Tools, Inc.) was used to cut and remove the cuticle above the brain. Connective tissue and the trachea above the brain were also removed. The petri dish was then filled with fly Ringer’s solution (103 mM NaCl, 3 mM KCl, 5 mM TES, 10 mM trehalose, 10 mM glucose, 9 mM sucrose, 26 mM NaHCO3, 1 mM NaH2PO4, 1.5 mM CaCl2, and 4 mM MgCl2, adjusted to 275–280 mOsm, bubbled with 95% O2/5%CO2) for acoustic coupling. Note that both during the preparation and during olfactory stimulation experiments, the antennae remained completely dry. A constant airflow (carrier stream; 0.8  L/min) was maintained across the fly antenna. For odor stimulation, a constant volume of headspace (0.5  L/min) above an odor bottle containing ethyl acetate was injected into the carrier stream.

2.4.

PACT of Acute Mouse Brain Slices

An optical parametric oscillator (OPO) laser (basiScan, Spectra-Physics, Inc.), pumped by an Nd: YAG laser (Qsmart 850, Quantel, Inc.), provided 488-nm-wavelength pulses with a 10-Hz pulse repetition rate. The laser beam was homogenized by an optical diffuser (EDC-5, RPC Photonics, Inc.). The PA signals were detected by a full-ring ultrasonic transducer array (Imasonic, Inc.) with a 10-cm diameter, 5-MHz central frequency, >90% one-way bandwidth, and 512 elements. Each element (20-mm height, 0.61-mm pitch, and 0.1-mm interelement space) was cylindrically focused to produce an axial focal length of 40 mm (acoustic NA: 0.25). The combined foci of all 512 elements formed a field of view of 20  mm diameter with an approximately isotropic in-plane resolution of 100  μm and a elevational resolution of 1  mm. The wide-field imaging exposure time was under 20  μs with a frame rate of 10 Hz. The raw data from each element was first Wiener deconvolved to account for the ultrasonic transducer’s impulse response and then reconstructed based on the dual speed-of-sound half-time backprojection algorithm to minimize acoustic heterogeneity.6163

A brain slice holder made of 4% agarose was placed in the center of the imaging chamber of the PACT system. A silicone tube was used for high-potassium perfusion (150 mM KCl concentration) in the imaging chamber. During imaging, the chamber was filled with a regular artificial cerebrospinal fluid (aCSF) solution and continuously bubbled with 5% CO2 and 95% O2. A peristaltic pump circulated and homogenized the aCSF solution within the imaging chamber.

2.5.

Brain Slice Preparation

The protocol for mouse brain slice preparation was approved by the Animal Studies Committee at Washington University in St. Louis. Acute coronal brain slices were prepared from 3 months old GCaMP6f (C57BL/6J-Tg) mice, which were purchased from Jackson Laboratory. The GCaMP6f labeled mice were anesthetized with isoflurane and decapitated. The brain was removed and glued onto a vibratome (Leica VT1200s, Leica Microsystems, Inc.) specimen holder. Coronal slices (300-μm thick) were cut in an ice-cold modified aCSF solution (125-mM NaCl, 25-mM glucose, 25-mM NaHCO3, 2.5-mM KCl, 1.25-mM NaH2PO4, 0.5-mM CaCl2, and 3-mM MgCl2) equilibrated with 5% CO2 and 95% O2. Slices were then transferred to a choline-based aCSF solution (92-mM choline chloride, 2.5-mM KCl, 1.2-mM NaH2PO4, 30-mM NaHCO3, 20-mM HEPES, 25-mM glucose, 5-mM Na-ascorbate, 2-mM thiourea, 3-mM Na-pyruvate, 2-mM CaCl2, and 1-mM MgCl2) equilibrated with 5% CO2 and 95% O2, incubated at 32°C for 30 min, subsequently stored in a regular aCSF solution (125-mM NaCl, 25-mM glucose, 25-mM NaHCO3, 2.5-mM KCl, 1.25-mM NaH2PO4, 2-mM CaCl2, and 1-mM MgCl2) equilibrated with 5% CO2 and 95% O2 at room temperature for more than an hour, and imaged with the PACT system.

2.6.

Wide-field Fluorescence Imaging of the Fly Brain and Acute Mouse Brain Slices

Wide-field fluorescence imaging was performed on an Olympus Fluoview 1000 system (excitation wavelength at 488 nm, emission filter wavelength centered at 510 nm; Olympus, Corp.). The fluorescence images were captured by an EMCCD camera (iXon Ultra 888, Andor, Ltd.). A 10× water immersion objective (0.3 NA, UMPLFLN, Olympus, Corp.) was used for GCaMP5G-expressing fly brain imaging. A 2× objective (0.14 NA, XLFluor2X/340, Olympus, Corp.) was used for GCaMP6f-expressing mouse brain slice imaging.

2.7.

Photoacoustic Microscopy of GCaMP Mouse Brain during Hindpaw Stimulations

To further validate our technique for in vivo GCaMP imaging in mouse brains, we used the same PAM system as described for fruit fly imaging. Anesthesia was induced by ketamine (100  mg/mL) and xylazine (20  mg/mL) and maintained with 0.5% to 1% isoflurane in oxygen, with ventilator support and supplemental temperature support. The scalp of the mouse brain was cut open to expose the skull for photoacoustic imaging and was surgically sutured after the experiment. During the imaging, the mouse was secured using a tooth bar on a platform. Forepaw stimulations were induced by a needle electrode inserted under the skin of the left forepaw. The electrode was connected to a function generator (DS345, Stanford Research Systems) for generating stimulation pulses. All of the animal experimental procedures were in conformance with the laboratory animal protocol approved by Washington University in St. Louis.

We first acquired the vasculature map of the mouse brain expressing GCaMP6s and then monitored the PA response of the mouse brain at the selected point (marked by a white cross) during left hindpaw stimulation trials. Each stimulation trial consisted of 5 s of pulsed stimulation (300-μs pulse width and 1-mA amplitude at 3 Hz). The resting period after each stimulation period was 20 s. The PA signal change during stimulations was averaged over seven stimulation trials and then lowpass filtered with a cutoff frequency of 5 Hz.

2.8.

Data Analysis

Image processing was performed in MATLAB (Mathworks, Inc.). In all experiments, the baseline PA and fluorescence signals were calculated as the averaged signals acquired before the simulation. The fractional changes in PA and fluorescence signals, compared with the baseline values, were then calculated.

3.

Results

3.1.

PAM of GCaMP5G-expressing Fruit Fly Brains in Vivo

One challenge for imaging GCaMP in live brains with PAT is the interference signals from hemoglobin, which has a strong optical absorption at the peak wavelength of the GCaMP absorption spectrum (488 nm). To overcome this challenge64 and unequivocally demonstrate the feasibility of GCaMP imaging with PAT, we chose a simple and well-established transgenic invertebrate model, Drosophila melanogaster (i.e., fruit fly), which does not have hemoglobin in its brain. For the proof-of-concept experiments, we used a transgenic fly line that expresses GCaMP5G pan-neutrally in all cholinergic neurons.

To image neural calcium signals in the fly brain, we developed a PA and epi-fluorescence dual-modality microscope [Fig. 1(a)]. The dual-modality microscope simultaneously recorded the odor-evoked neural activities in the fly brain in vivo, with an automatically coregistered field of view and complementary image contrasts. When a 488-nm–wavelength laser pulse excites the GCaMP molecules in the brain, the captured laser energy is transformed partially into fluorescence signals and partially into PA signals, with a 67% fluorescence quantum yield.65 It is worth noting that the ultrasonic transducer in our microscope temporally resolves PA signals from different depths along the acoustic axis, and no additional depth-scanning is required. Moreover, only a 2D raster scan is needed for PAM to form a 3D image. The PAM system and the epifluorescence microscopy (EFM) system share the same optical focusing pathway and thus have the same lateral resolution. The axial resolution of PAM is acoustically determined, whereas EFM lacks axial resolution over planar targets.

Using PAM, we first imaged the whole fly brain with the cuticle intact as well as the adjacent body parts. A maximum amplitude projection (MAP) image clearly shows the brain cuticle, antennae, compound eyes, and thorax [Fig. 1(b)]. The strong PA signals mostly come from the pigments of the cuticle. To gain direct optical access to neurons in the fly brain, we removed the cuticle posterior to the two antennae and the underlying connective tissue surrounding the brain.59 The PA depth-resolved images of the exposed fly brain show heterogeneous GCaMP expression levels in the brain [Fig. S1(a) in the Supplemental Material]. The strongest PA signals came from putative neural cell bodies. To confirm that the PA signal was generated by GCaMP, we used PAM to measure the absorption spectrum of the brain tissue by tuning the laser wavelength from 410 to 550 nm [Fig. S1(b) in the Supplemental Material]. The spectrum measured by PAM agrees well with that measured purely optically elsewhere.65 This result is strong evidence that GCaMP is the major chromophore in the fly brain in the visible spectral range. In addition, fluorescence signals from GCaMP were simultaneously acquired by the EFM channel [Fig. 2(a)] to further confirm the origin of the PA signals.

Fig. 2

PA imaging of odor-evoked neural activity in the fly brain without and with cuticle. (a) Epi-fluorescence image of a GCaMP5G-expressing fly brain with the cuticle surgically removed. The dashed circle identifies the lateral location of M-mode PA recording. (b) Fractional PA amplitude changes from representative neuron layers in the glomerular region of the fly antennal lobe, showing varied response amplitudes and durations at different depths. Note that the depth in this panel is measured from the surface of the fly brain. The gray bar indicates when a puff of ethyl acetate was delivered to the fly antenna. (c) Averaged fractional PA amplitude response and the concurrently recorded fluorescence intensity response to odor stimulation with the cuticle removed. (d) Epifluorescence image of a GCaMP5G-expressing fly brain with the cuticle intact (i.e., not surgically removed). The lateral location of the M-mode PA recording is marked by the dashed circle. (e) Fractional PA amplitude changes from representative neuron layers in the fly brain with its cuticle intact. Note that the lack of response in the cuticle layer (CL) is shown here as a control. Depth in this panel is measured from the cuticle layer. The gray bar indicates the duration of the ethyl acetate delivery to the fly antennae. (f) Averaged fractional PA amplitude response from the fly brain with intact cuticle and the concurrently recorded fluorescence intensity response to odor stimulation. Although the fractional changes in PA amplitude and fluorescence intensity are comparable in magnitude, depth-resolved measurements of neural activity were made only by the PA system. (Video 1, MP4, 286 KB [URL: https://doi.org/10.1117/1.JBO.27.9.096004.s1]; Video 2, MP4, 640 KB [URL: https://doi.org/10.1117/1.JBO.27.9.096004.s2)

JBO_27_9_096004_f002.png

We monitored neural activities in the fly brain evoked by a sensory stimulus. We administered a 5-s long puff of ethyl acetate to the fruit fly antenna while monitoring the changes in the GCaMP signals in the glomerular regions of the antennal lobe (a region where olfactory sensory neuron axon terminals converge and synapse with the downstream neural populations). M-mode recording (i.e., repeated 1D depth-resolved PA imaging at the same lateral location) was performed in the fruit fly antennal lobe. Depth-resolved PA signals from different layers in the brain (200-μm thick) were acquired at a 1D imaging rate of 1 kHz [Fig. 2(b)]. The increase in GCaMP’s optical absorption coefficient led to the same fractional PA and fluorescent signal changes. In response to the odor stimulation, an 23% increase was observed in depth-averaged PA signals [Fig. 2(c)], which indicated elevated neural activities (Video 1). The strongest fractional PA signal change (34%) was observed at 90-μm beneath the brain surface. We observed different response durations at different depths from the PAM channel, which clearly indicates the heterogeneity in the dynamics of odor-evoked responses from sensory neuron axonal terminals. Ethyl acetate strongly activates different types of sensory neurons, particularly those housed in the large basiconic sensilla. The axonal terminals of these sensory neurons are known arborize in different glomeruli at different depths in the fly antennal lobe.66,67 The PA response time course averaged over all depths correlates well with the fluorescence response [Fig. 2(c)] simultaneously recorded at the same lateral location. This correlation indicates that the same neuronal events were being imaged by both the PAM and EFM channels. More importantly, depth-resolved signals were only obtained using PAM, clearly illustrating the benefit of the proposed approach. We also recorded similar PA and fluorescence responses in the lateral horn region of the fly brain (Fig. S2 in the Supplemental Material).

Next, taking advantage of the depth-resolving and scattering-tolerant capabilities of PAM, we sought to determine whether such depth-resolved signals can be obtained from flies with intact cuticles (i.e., without any surgical preparations) [Fig. 2(d)]. Although light was attenuated by the cuticle and connective tissue, PAM still detected signals from all neuron layers in the brain. We recorded concurrent M-mode PA and fluorescence signals upon odor stimulation [Fig. 2(e), Video 2). Similar to the results acquired with the cuticle removed, different PA responses were recorded from different neuronal layers [Fig. 2(e)]. The PA signals from the surface layer of the cuticle, serving as a built-in control, did not show any changes during odor stimulation, as expected. Note that, even with the intact cuticle, the fractional signal change was observed reliably [Figs. 2(e) and 2(f), Fig. S3 in the Supplemental Material]. Again, the PA results were confirmed by the concurrent fluorescence recording [Fig. 2(f)].

As mentioned earlier, PAM can render a 3D calcium image with only a 2D scan. To illustrate this, we imaged a volume of 100  μm (x axis) by 40  μm (y axis) by 270  μm (z axis) in the antennal lobe region of the fly brain (with the cuticle removed), at a volumetric imaging rate of 1 Hz (Fig. 3, Fig. S4 in the Supplemental Material, and Videos 3 and 4). Both anatomical structures and the functional activity of neurons expressing GCaMP5G in the putative glomerular layers of the fly antennal lobe were clearly imaged [Fig. 3(a)]. The diversity of odor-evoked responses at different depths was still captured [Figs. 3(b) and 3(c)]. In sum, these results convincingly show that PAM can monitor depth-resolved calcium signals from neurons with high spatial and temporal resolutions.

Fig. 3

Volumetric PAM of odor-evoked neural activities in the glomerular antennal lobe region of a fly antennal lobe. (a) Volumetric images of the fly brain with the cuticle removed, before, during, and after two odor stimulations. The fractional changes in PA amplitude (shown in color) are overlaid on the structural image (shown in gray). (b) Depth-wise averaged fractional changes in PA amplitude during odor stimulations. The changes are averaged over the two stimulations. (c) Time courses of averaged fractional PA amplitude changes from two representative layers [labeled in (a)] of neurons. The gray bars indicate two successive puffs of ethyl acetate. (Video 3, MP4, 502 KB [URL: https://doi.org/10.1117/1.JBO.27.9.096004.s3]; Video 4, MP4, 392 KB [URL: https://doi.org/10.1117/1.JBO.27.9.096004.s4])

JBO_27_9_096004_f003.png

3.2.

PACT of Live GCaMP6f-Expressing Mouse Brain Slices ex Vivo

Taking advantage of PACT’s deep imaging capability, we used our recently developed high-speed PACT system to image neural activity at depths greater than the optical diffusion limit of 1  mm [Fig. 4(a)].68 With wide-field illumination and parallel acoustic detection, the newly developed PACT system features high spatiotemporal resolution, deep penetration, and full-view fidelity. PACT acquires a 2D cross-sectional image with a 100-μm resolution at a 10-Hz frame rate. Further, only a single laser pulse is used to generate each 2D image, thereby greatly reducing unnecessary light exposure and resultant photobleaching. In the absence of hemodynamically induced PA signal changes in the brain slice, GCaMP6f is the chromophore that can generate dynamic PA signals upon 488-nm laser excitation during neural activities.

Fig. 4

PACT monitoring of in vitro neural activities in GCaMP6f-expressing mouse brain slices. (a) Schematic of the high-speed PACT system, with wide-field illumination and parallel acoustic detection. The GCaMP6f-expressing brain slice was immersed in a regular aCSF solution in the imaging chamber. (b) PACT images of the GCaMP6f-expressing brain slice in a clear medium (left) and through a 2-mm-thick scattering Intralipid solution (right, reduced optical scattering coefficient 10  cm1). (c) PACT of the neural activity in a GCaMP6f-expressing mouse brain slice in response to high-potassium perfusion. The fractional PA amplitude change (shown in color) is superimposed on the baseline image (shown in gray). (d) Time courses of fractional PA amplitude changes from four representative regions (I, II, III, and IV) as indicated in (c). The solid lines are fittings using a bell-shaped dose-response model. (Video 5, MP4, 2046 KB [URL: https://doi.org/10.1117/1.JBO.27.9.096004.s5]; Video 6, MP4, 265 KB [URL: https://doi.org/10.1117/1.JBO.27.9.096004.s6])

JBO_27_9_096004_f004.png

We imaged neural activity in acute coronal slices (300-μm thick) of a GCaMP6f-expressing mouse brain. The brain slice was mounted in the PACT imaging chamber, which was filled with regular aCSF, and imaged either without any scattering medium or through a 2-mm-thick tissue-simulating scattering medium [Fig. 4(b)]. In PACT images acquired without an overlying scattering medium, major brain structures expressing GCaMP6f, such as the cortex and hippocampus, were clearly resolved. Notably, even when the brain slice was covered with a 2-mm-thick scattering medium (1% Intralipid in deionized water, with a reduced scattering coefficient 10  cm1), GCaMP6f-rich brain structures were still observable [Fig. 4(b)].

To stimulate the neurons in the brain slice, we performed high-concentration potassium perfusion in the PACT imaging chamber. The elevated extracellular potassium depolarized neurons in the brain slice and induced action potentials. Upon high-potassium perfusion, increased PA signals were recorded from multiple areas within the brain slice [Fig. 4(c), Video 5]. The fractional PA amplitude changes shown in Fig. 4(d) were averaged over representative regions of interest. The fractional PA amplitude increased by up to 15% during the potassium perfusion. After the potassium perfusion, the intracellular calcium concentration gradually returned to the baseline level through diffusion and sodium-calcium exchanges, resulting in decreased PA amplitudes in the originally responding areas. In another experiment with a mouse brain slice covered with 1.5-mm-thick mouse brain tissue, we also observed a similar PA signal increase upon high-potassium perfusion. The PA response was independently validated by monitoring the fluorescence response of a brain slice in regular aCSF, measured with a commercial wide-field fluorescence microscope. The same sample preparation and stimulation procedures were used (Fig. S5 in the Supplemental Material and Video 6). In sum, our results confirm that PACT is a unique and effective calcium imaging technique with an image depth beyond the optical diffusion limit.

3.3.

PAM of a GCaMP6s-Expressing Mouse Brain in Vivo

In addition to optical scattering, another obstacle to photoacoustic neural imaging in mammalian brains is the presence of strong hemodynamic signals, which may overshadow the changes in GCaMP signals. By imaging spatially confined areas with minimal vasculature, the contribution of hemodynamics can be minimized. In our pilot study, we first acquired the vasculature map of a mouse brain expressing GCaMP6s [Fig. S6(a) in the Supplemental Material]. Then we monitored the PA signal change of a selected point with minimal vasculature over the stimulation trials. It can be seen that the local PA response from the mouse brain in vivo shows a relatively slow increase over the 5-s stimulation period [Fig. S6(b) in the Supplemental Material] with 15 stimulation pulses at 3 Hz. Interestingly, after the start of each stimulation pulse, a fast increase in PA signal can be observed on top of the slow signal change. Considering the relatively fast temporal response of GCaMP compared with that of hemodynamic signals, we believe that the fast increase in PA signals came from GCaMP and the slow change came from the hemodynamic change.

4.

Discussion

We have demonstrated, for the first time, that multiscale PAT can be used for imaging neuronal calcium signals, with high spatiotemporal resolution and deep penetration. With fast dynamics, high photostability, and a large fractional absorption response, GCaMP or its variants offers an ideal optical contrast for PAT as the PA signal is exquisitely sensitive to optical absorption. With the depth-resolved imaging capability, we showed that PAM can perform fast 3D calcium imaging in fruit flies expressing genetically encoded calcium indicators. Notably, using PAM, we were able to monitor depth-resolved odor-evoked neural activities from the fly brain, without the need for invasive surgery, which could compromise experimental reproducibility.69 For in vivo mouse brain imaging, we showed that GCaMP signals can be temporally separated from hemodynamic signals. To unambiguously differentiate hemodynamic signals and GCaMP signals in vivo, multiwavelength PAT can be implemented to simultaneously record hemodynamic and GCaMP signals, enabling studies of neurovascular coupling. Although we focused on the feasibility of PAM for neural imaging here, the imaging rate can be readily increased by utilizing a higher pulse-repetition-rate laser and an optimized scanning mechanism.7072

Optical scattering of light severely limits high-resolution optical imaging techniques to superficial tissue layers. Our results clearly demonstrate that, by utilizing diffused photons, PACT can image neural responses beyond the optical diffusion limit. Prior work has shown that PACT can provide a penetration depth of 8 mm in a rat brain.68 Although, the imaging depth of PACT demonstrated in this study was largely limited by the relatively short absorption wavelength of GCaMP, it can be further improved using calcium indicators that absorb strongly at longer wavelengths,52,73,74 especially in the near-infrared wavelength region (700 to 1100 nm). The combination of near-infrared calcium indicators and PACT can potentially enable small-animal whole-brain neural imaging.

So far, a variety of imaging modalities have been explored for brain studies, especially two-photon microscopy and functional magnetic resonance imaging (fMRI). At the microscopic scale, two-photon microscopy allows for detailed and direct examination of neural activities in small animal models.75 At the macroscopic scale, fMRI has been extensively used to map human or large-animal brain activities, allowing for indirect coarse-grained localization of cognitive functions with hemodynamic signals. However, correlating the two imaging scales still remains a major challenge due to the different signal-contrast sources of these two modalities. Therefore, there is a gap in our understanding between the macroscopic activity patterns in humans or large animals and the microscopic activity details in small-animal models. PAT has been scaled up for multiple-centimeter penetration and can provide multiscale imaging for not only small animals but also potentially large animals (at least with open skulls).76,77 Multiwavelength PAT can be implemented to simultaneously record and unmix hemodynamic and calcium signals, enabling studies of neurovascular coupling. Therefore, PAT is uniquely positioned to bridge this contrast gap across the length scales, facilitating correlative neuroscience studies in various animal models of brain diseases such as Alzheimer’s disease and stroke.7

We have shown that PAT holds great promise for neuronal calcium imaging in the brains of transgenic invertebrate and vertebrate models expressing GCaMP indicators. With PAM, we demonstrate in vivo high-resolution depth-resolved neural imaging without depth scanning that enables 3D imaging of neural activities at high temporal resolution. With PACT, we demonstrate neural activity imaging with GCaMP beyond the optical diffusion limit. The combination of advanced PAT technologies and genetically encoded neural activity indicators optimized for PA imaging offers the neuroscience community a methodology for studying neural networks in vivo.

5.

Conclusion

We have shown that PAT holds great promise for neuronal calcium imaging in the brains of transgenic invertebrate and vertebrate models expressing GCaMP indicators. With PAM, we demonstrate in vivo high-resolution depth-resolved neural imaging without depth scanning that enables 3D imaging of neural activities at high temporal resolution. With PACT, we demonstrate neural activity imaging with GCaMP beyond the optical diffusion limit. The combination of advanced PAT technologies and genetically encoded neural activity indicators optimized for PA imaging offers the neuroscience community a methodology for studying neural networks in vivo.

Disclosure

L.V.W. has financial interests in Microphotoacoustics, Inc.; CalPACT, LLC; and Union Photoacoustic Technologies, Ltd., which did not support this work.

Acknowledgments

The authors thank Prof. James Ballard for a close reading of the manuscript and Terence Wong for experimental help. This work was sponsored in part by the US National Institutes of Health (NIH), grants U01 NS099717 (BRAIN Initiative) to Lihong Wang and 5K99AR062530 (NIH K99/R00) to Bin Rao; an Imaging Sciences Pathway fellowship to Haoyang Rong; and a National Science Foundation (NSF) CAREER grant (#1453022) to Baranidharan Raman.

Author Contributions

R. Z., Bi. R., Ba. R., and L. V. W. conceived and designed the study. R. Z., L. L., and Bi. R. constructed the imaging systems. R. Z., L. L., Bi. R., and J. Y. performed the PA experiments and analyzed the data. H. R. performed the drosophila surgeries. M. S. and S. M. helped with brain slice preparation. R. C. and Q. Z. fabricated the ultrasonic transducer. Ba. R. and L. V. W. supervised the study. All authors discussed the results and wrote the manuscript.

References

1. 

L. L. Looger and O. Griesbeck, “Genetically encoded neural activity indicators,” Curr. Opin. Neurobiol., 22 18 –23 (2012). https://doi.org/10.1016/j.conb.2011.10.024 COPUEN 0959-4388 Google Scholar

2. 

A. Devor et al., “The challenge of connecting the dots in the BRAIN,” Neuron, 80 270 –274 (2013). https://doi.org/10.1016/j.neuron.2013.09.008 NERNET 0896-6273 Google Scholar

3. 

M. E. Raichle and M. A. Mintun, “Brain work and brain imaging,” Annu. Rev. Neurosci., 29 449 –476 (2006). https://doi.org/10.1146/annurev.neuro.29.051605.112819 ARNSD5 0147-006X Google Scholar

4. 

T. A. Christensen et al., “Multi-unit recordings reveal context-dependent modulation of synchrony in odor-specific neural ensembles,” Nat. Neurosci., 3 927 –931 (2000). https://doi.org/10.1038/78840 NANEFN 1097-6256 Google Scholar

5. 

R. Eckhorn and U. Thomas, “A new method for the insertion of multiple microprobes into neural and muscular tissue, including fiber electrodes, fine wires, needles and microsensors,” J. Neurosci. Methods, 49 175 –179 (1993). https://doi.org/10.1016/0165-0270(93)90121-7 JNMEDT 0165-0270 Google Scholar

6. 

D. Saha et al., “Multi-unit recording methods to characterize neural activity in the locust (Schistocerca Americana) olfactory circuits,” J. Visual. Exp., e50139 (2013). Google Scholar

7. 

E. M. C. Hillman, “Optical brain imaging in vivo: techniques and applications from animal to man,” J. Biomed. Opt., 12 051402 (2007). https://doi.org/10.1117/1.2789693 JBOPFO 1083-3668 Google Scholar

8. 

B. Kuhn, W. Denk and R. M. Bruno, “In vivo two-photon voltage-sensitive dye imaging reveals top-down control of cortical layers 1 and 2 during wakefulness,” Proc. Natl. Acad. Sci. U. S. A., 105 (21), 7588 –7593 (2008). https://doi.org/10.1073/pnas.0802462105 Google Scholar

9. 

J. N. D. Kerr and W. Denk, “Imaging in vivo: watching the brain in action,” Nat. Rev. Neurosci., 9 195 –205 (2008). https://doi.org/10.1038/nrn2338 NRNAAN 1471-003X Google Scholar

10. 

M. Scanziani and M. Hausser, “Electrophysiology in the age of light,” Nature, 461 930 –939 (2009). https://doi.org/10.1038/nature08540 Google Scholar

11. 

R. Cossart, D. Aronov and R. Yuste, “Attractor dynamics of network UP states in the neocortex,” Nature, 423 283 –288 (2003). https://doi.org/10.1038/nature01614 Google Scholar

12. 

R. Yuste and L. C. Katz, “Control of postsynaptic Ca2+ influx in developing neocortex by excitatory and inhibitory neurotransmitters,” Neuron, 6 333 –344 (1991). https://doi.org/10.1016/0896-6273(91)90243-S NERNET 0896-6273 Google Scholar

13. 

R. Amor et al., “Widefield two-photon excitation without scanning: live cell microscopy with high time resolution and low photo-bleaching,” PLoS One, 11 e0147115 (2016). https://doi.org/10.1371/journal.pone.0147115 POLNCL 1932-6203 Google Scholar

14. 

E. Y. Yew, C. J. Sheppard and P. T. So, “Temporally focused wide-field two-photon microscopy: paraxial to vectorial,” Opt. Express, 21 12951 –12963 (2013). https://doi.org/10.1364/OE.21.012951 OPEXFF 1094-4087 Google Scholar

15. 

S. Takanezawa, T. Saitou and T. Imamura, “Wide field light-sheet microscopy with lens-axicon controlled two-photon Bessel beam illumination,” Nat. Commun., 12 2979 (2021). https://doi.org/10.1038/s41467-021-23249-y NCAOBW 2041-1723 Google Scholar

16. 

T. A. Planchon et al., “Rapid three-dimensional isotropic imaging of living cells using Bessel beam plane illumination,” Nat. Methods, 8 417 –423 (2011). https://doi.org/10.1038/nmeth.1586 1548-7091 Google Scholar

17. 

L. Gao et al., “3D live fluorescence imaging of cellular dynamics using Bessel beam plane illumination microscopy,” Nat. Protoc., 9 1083 –1101 (2014). https://doi.org/10.1038/nprot.2014.087 1754-2189 Google Scholar

18. 

A. F. Valle and J. D. Seelig, “Two-photon Bessel beam tomography for fast volume imaging,” Opt. Express, 27 12147 –12162 (2019). https://doi.org/10.1364/OE.27.012147 OPEXFF 1094-4087 Google Scholar

19. 

T. Wang and C. Xu, “Three-photon neuronal imaging in deep mouse brain,” Optica, 7 947 –960 (2020). https://doi.org/10.1364/OPTICA.395825 Google Scholar

20. 

D. G. Ouzounov et al., “In vivo three-photon imaging of activity of GCaMP6-labeled neurons deep in intact mouse brain,” Nat. Methods, 14 388 –390 (2017). https://doi.org/10.1038/nmeth.4183 1548-7091 Google Scholar

21. 

D. M. Chow et al., “Deep three-photon imaging of the brain in intact adult zebrafish,” Nat. Methods, 17 605 –608 (2020). https://doi.org/10.1038/s41592-020-0819-7 1548-7091 Google Scholar

22. 

N. G. Horton et al., “In vivo three-photon microscopy of subcortical structures within an intact mouse brain,” Nat. Photonics, 7 205 –209 (2013). https://doi.org/10.1038/nphoton.2012.336 NPAHBY 1749-4885 Google Scholar

23. 

Y. Y. Gong et al., “High-speed recording of neural spikes in awake mice and flies with a fluorescent voltage sensor,” Science, 350 1361 –1366 (2015). https://doi.org/10.1126/science.aab0810 SCIEAS 0036-8075 Google Scholar

24. 

D. G. Ouzounov et al., “In vivo three-photon calcium imaging of brain activity from layer 6 neurons in mouse brain,” in Conf. Lasers and Electro-Opt. (CLEO) - Laser Sci. to Photonic Appl., 1 –2 (2014). https://doi.org/10.1364/CLEO_SI.2014.STh5C.2 Google Scholar

25. 

L. Li et al., “Snapshot photoacoustic topography through an ergodic relay of optical absorption in vivo,” Nat. Protoc., 16 2381 –2394 (2021). https://doi.org/10.1038/s41596-020-00487-w 1754-2189 Google Scholar

26. 

L. Li, J. Yao, L. V. Wang, “Photoacoustic tomography of neural systems,” Neural Engineering, 349 –378 Springer International Publishing, Cham (2020). Google Scholar

27. 

B. A. Flusberg et al., “High-speed, miniaturized fluorescence microscopy in freely moving mice,” Nat. Methods, 5 935 –938 (2008). https://doi.org/10.1038/nmeth.1256 1548-7091 Google Scholar

28. 

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

29. 

T. Berdyyeva et al., “Zolpidem reduces hippocampal neuronal activity in freely behaving mice: a large scale calcium imaging study with miniaturized fluorescence microscope,” PloS One, 9 e112068 (2014). https://doi.org/10.1371/journal.pone.0112068 POLNCL 1932-6203 Google Scholar

30. 

L. A. Gunaydin et al., “Natural neural projection dynamics underlying social behavior,” Cell, 157 1535 –1551 (2014). https://doi.org/10.1016/j.cell.2014.05.017 CELLB5 0092-8674 Google Scholar

31. 

L. Li et al., “Small near-infrared photochromic protein for photoacoustic multi-contrast imaging and detection of protein interactions in vivo,” Nat. Commun., 9 2734 (2018). https://doi.org/10.1038/s41467-018-05231-3 NCAOBW 2041-1723 Google Scholar

32. 

D. Razansky et al., “Multispectral opto-acoustic tomography of deep-seated fluorescent proteins in vivo,” Nat. Photonics, 3 412 –417 (2009). https://doi.org/10.1038/nphoton.2009.98 NPAHBY 1749-4885 Google Scholar

33. 

J. J. Yao et al., “Multiscale photoacoustic tomography using reversibly switchable bacterial phytochrome as a near-infrared photochromic probe,” Nat. Methods, 13 67 –73 (2016). https://doi.org/10.1038/nmeth.3656 1548-7091 Google Scholar

34. 

Z. Wu et al., “A microrobotic system guided by photoacoustic computed tomography for targeted navigation in intestines in vivo,” Sci. Rob., 4 eaax0613 (2019). Google Scholar

35. 

L. Li and L. V. Wang, “Recent advances in photoacoustic tomography,” BME Front., 2021 1 –17 (2021). https://doi.org/10.34133/2021/9823268 Google Scholar

36. 

C. Kim, C. Favazza and L. H. V. Wang, “In vivo photoacoustic tomography of chemicals: high-resolution functional and molecular optical imaging at new depths,” Chem. Rev., 110 2756 –2782 (2010). https://doi.org/10.1021/cr900266s CHREAY 0009-2665 Google Scholar

37. 

B. Rao et al., “Photoacoustic imaging of voltage responses beyond the optical diffusion limit,” Sci. Rep., 7 2560 (2017). https://doi.org/10.1038/s41598-017-02458-w SRCEC3 2045-2322 Google Scholar

38. 

L. Li et al., “Integration of multitargeted polymer-based contrast agents with photoacoustic computed tomography: an imaging technique to visualize breast cancer intratumor heterogeneity,” ACS Nano, 15 2413 –2427 (2021). https://doi.org/10.1021/acsnano.0c05893 ANCAC3 1936-0851 Google Scholar

39. 

L. Li et al., “Multiscale photoacoustic tomography of a genetically encoded near-infrared FRET biosensor,” Adv. Sci., 8 2102474 (2021). https://doi.org/10.1002/advs.202102474 Google Scholar

40. 

Y. Qu et al., “Dichroism-sensitive photoacoustic computed tomography,” Optica, 5 495 –501 (2018). https://doi.org/10.1364/OPTICA.5.000495 Google Scholar

41. 

X. D. Wang et al., “Noninvasive laser-induced photoacoustic tomography for structural and functional in vivo imaging of the brain,” Nat. Biotechnol., 21 803 –806 (2003). https://doi.org/10.1038/nbt839 NABIF9 1087-0156 Google Scholar

42. 

Z. L. Deng et al., “In vivo imaging of hemodynamics and oxygen metabolism in acute focal cerebral ischemic rats with laser speckle imaging and functional photoacoustic microscopy,” J. Biomed. Opt., 17 081415 (2012). https://doi.org/10.1117/1.JBO.17.8.081415 JBOPFO 1083-3668 Google Scholar

43. 

L. Li et al., “Fully motorized optical-resolution photoacoustic microscopy,” Opt. Lett., 39 2117 –2120 (2014). https://doi.org/10.1364/OL.39.002117 OPLEDP 0146-9592 Google Scholar

44. 

J. Jo and X. Yang, “Functional photoacoustic imaging to observe regional brain activation induced by cocaine hydrochloride,” J. Biomed. Opt., 16 090506 (2011). https://doi.org/10.1117/1.3626576 JBOPFO 1083-3668 Google Scholar

45. 

L. Li et al., “Label-free photoacoustic tomography of whole mouse brain structures ex vivo,” Neurophotonics, 3 035001 (2016). https://doi.org/10.1117/1.NPh.3.3.035001 Google Scholar

46. 

Y. Y. Liu et al., “Assessing the effects of norepinephrine on single cerebral microvessels using optical-resolution photoacoustic microscope,” J. Biomed. Opt., 18 076007 (2013). https://doi.org/10.1117/1.JBO.18.7.076007 JBOPFO 1083-3668 Google Scholar

47. 

P. Zhang et al., “High‐resolution deep functional imaging of the whole mouse brain by photoacoustic computed tomography in vivo,” J. Biophotonics, 11 e201700024 (2018). https://doi.org/10.1002/jbio.201700024 Google Scholar

48. 

M. C. Pilatou et al., “Photoacoustic imaging of brain perfusion on albino rats by using Evans blue as contrast agent,” Arch. Physiol. Biochem., 111 389 –397 (2003). https://doi.org/10.3109/13813450312331337649 Google Scholar

49. 

V. Tsytsarev et al., “Photoacoustic and optical coherence tomography of epilepsy with high temporal and spatial resolution and dual optical contrasts,” J. Neurosci. Methods, 216 142 –145 (2013). https://doi.org/10.1016/j.jneumeth.2013.04.001 JNMEDT 0165-0270 Google Scholar

50. 

L. D. Wang, K. Maslov and L. H. V. Wang, “Single-cell label-free photoacoustic flowoxigraphy in vivo,” Proc. Natl. Acad. Sci. U. S. A., 110 (15), 5759 –5764 (2013). https://doi.org/10.1073/pnas.121557811 Google Scholar

51. 

P. O’Herron et al., “Neural correlates of single-vessel haemodynamic responses in vivo,” Nature, 534 378 –382 (2016). https://doi.org/10.1038/nature17965 Google Scholar

52. 

J. Akerboom et al., “Genetically encoded calcium indicators for multi-color neural activity imaging and combination with optogenetics,” Front. Mol. Neurosci., 6 2 (2013). https://doi.org/10.3389/fnmol.2013.00002 Google Scholar

53. 

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

54. 

S. Gottschalk et al., “Rapid volumetric optoacoustic imaging of neural dynamics across the mouse brain,” Nat. Biomed. Eng., 3 392 –401 (2019). https://doi.org/10.1038/s41551-019-0372-9 Google Scholar

55. 

L. Tian et al., “Imaging neural activity in worms, flies and mice with improved GCaMP calcium indicators,” Nat. Methods, 6 875 –881 (2009). https://doi.org/10.1038/nmeth.1398 1548-7091 Google Scholar

56. 

J. W. Wang et al., “Two-photon calcium imaging reveals an odor-evoked map of activity in the fly brain,” Cell, 112 271 –282 (2003). https://doi.org/10.1016/S0092-8674(03)00004-7 CELLB5 0092-8674 Google Scholar

57. 

J. Mütze et al., “Excitation spectra and brightness optimization of two-photon excited probes,” Biophys. J., 102 934 –944 (2012). https://doi.org/10.1016/j.bpj.2011.12.056 BIOJAU 0006-3495 Google Scholar

58. 

L. D. Wang et al., “Video-rate functional photoacoustic microscopy at depths,” J. Biomed. Opt., 17 106007 (2012). https://doi.org/10.1117/1.JBO.17.10.106007 JBOPFO 1083-3668 Google Scholar

59. 

A. F. Silbering et al., “Calcium imaging of odor-evoked responses in the Drosophila antennal lobe,” J. Visual. Exp., (61), 2976 (2012). https://doi.org/10.3791/2976 Google Scholar

60. 

M. Murthy and G. Turner, “Dissection of the head cuticle and sheath of living flies for whole-cell patch-clamp recordings in the brain,” Cold Spring Harb. Protoc., 2013 134 –139 (2013). https://doi.org/10.1101/pdb.prot071696 Google Scholar

61. 

P. Hu et al., “Spatiotemporal antialiasing in photoacoustic computed tomography,” IEEE Trans. Med. Imaging, 39 3535 –3547 (2020). https://doi.org/10.1109/TMI.2020.2998509 ITMID4 0278-0062 Google Scholar

62. 

T. P. Matthews et al., “Parameterized joint reconstruction of the initial pressure and sound speed distributions for photoacoustic computed tomography,” SIAM J. Imaging Sci., 11 1560 –1588 (2018). https://doi.org/10.1137/17M1153649 Google Scholar

63. 

J. Poudel et al., “Mitigation of artifacts due to isolated acoustic heterogeneities in photoacoustic computed tomography using a variable data truncation-based reconstruction method,” J. Biomed. Opt., 22 041018 (2017). https://doi.org/10.1117/1.JBO.22.4.041018 JBOPFO 1083-3668 Google Scholar

64. 

B. Cox et al., “Quantitative spectroscopic photoacoustic imaging: a review,” J.Biomed. Opt., 17 061202 (2012). https://doi.org/10.1117/1.JBO.17.6.061202 JBOPFO 1083-3668 Google Scholar

65. 

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

66. 

A. Couto, M. Alenius and B. J. Dickson, “Molecular, anatomical, and functional organization of the Drosophila olfactory system,” Curr. Biol., 15 1535 –1547 (2005). https://doi.org/10.1016/j.cub.2005.07.034 CUBLE2 0960-9822 Google Scholar

67. 

A. A. Dobritsa et al., “Integrating the molecular and cellular basis of odor coding in the Drosophila antenna,” Neuron, 37 827 –841 (2003). https://doi.org/10.1016/S0896-6273(03)00094-1 NERNET 0896-6273 Google Scholar

68. 

L. Li et al., “Single-impulse panoramic photoacoustic computed tomography of small-animal whole-body dynamics at high spatiotemporal resolution,” Nat. Biomed. Eng., 1 0071 (2017). https://doi.org/10.1038/s41551-017-0071 Google Scholar

69. 

S. Sinha et al., “High-speed laser microsurgery of alert fruit flies for fluorescence imaging of neural activity,” Proc. Natl. Acad. Sci., 110 18374 –18379 (2013). https://doi.org/10.1073/pnas.1216287110 Google Scholar

70. 

J. J. Yao et al., “High-speed label-free functional photoacoustic microscopy of mouse brain in action,” Nat. Methods, 12 407 –410 (2015). https://doi.org/10.1038/nmeth.3336 1548-7091 Google Scholar

71. 

Y. He et al., “Wave of single-impulse-stimulated fast initial dip in single vessels of mouse brains imaged by high-speed functional photoacoustic microscopy,” J. Biomed. Opt., 25 066501 (2020). https://doi.org/10.1117/1.JBO.25.6.066501 JBOPFO 1083-3668 Google Scholar

72. 

Y. Li et al., “Snapshot photoacoustic topography through an ergodic relay for high-throughput imaging of optical absorption,” Nat. Photonics, 14 164 –170 (2020). https://doi.org/10.1038/s41566-019-0576-2 NPAHBY 1749-4885 Google Scholar

73. 

M. Inoue et al., “Rational design of a high-affinity, fast, red calcium indicator R-CaMP2,” Nat. Methods, 12 893 –893 (2015). https://doi.org/10.1038/nmeth0915-893b 1548-7091 Google Scholar

74. 

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

75. 

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

76. 

L. M. Nie et al., “Photoacoustic tomography through a whole adult human skull with a photon recycler,” J. Biomed. Opt., 17 110506 (2012). https://doi.org/10.1117/1.JBO.17.11.110506 JBOPFO 1083-3668 Google Scholar

77. 

S. Na et al., “Massively parallel functional photoacoustic computed tomography of the human brain,” Nat. Biomed. Eng., 6 (5), 584 –592 (2022). https://doi.org/10.1038/s41551-021-00735-8 Google Scholar

Biography

Ruiying Zhang received her BS degree in optical information science and technology from Wuhan University, Hubei, China, and her PhD from Washington University, St. Louis, in 2016. Her research interests include optical and photoacoustic/ultrasound imaging with biomedical applications.

Lei S. Li received his MS degree from Washington University, St. Louis, in 2016, and his PhD from California Institute of Technology in 2019. His current research focuses on technology development of photoacoustic tomography, especially improving the photoacoustic wholebody imaging performance and applying it on functional brain imaging and cancer detection.

Bin Rao received his PhD in biomedical optics from Beckman Laser Institute & Medical Clinic (BLI-MC), University of California, Irvine, in 2008. The ophthalmic Doppler optical coherence tomography system out of his PhD work is being used in a clinical study at BLI-MC. His postdoctoral trainings at Washington University, St. Louis, include photoacoustic microscopy, confocal, and two-photon microscopy. He was awarded the NIH Pathway to Independence Award in 2012.

Steven Mennerick received his PhD from Washington University, St. Louis, in 1995. He received postdoctoral training at State University of New York. He is now a professor of psychiatry at Washington University School of Medicine, St. Louis. The broad aim of his research is to understand control of neuronal excitation and inhibition by neurotransmitters in the central nervous system.

Baranidharan Raman is a professor of biomedical engineering at Washington University, St. Louis. Prior to joining Washington University, St. Louis, in February 2010, he was a joint postdoctoral fellow at the National Institutes of Health and the National Institute of Standards and Technology. His research interests include computational and systems neuroscience, pattern recognition, sensor-based machine olfaction, and bio-inspired intelligent systems.

Lihong V. Wang received his PhD from Rice University, Houston, Texas, United States. He currently holds the Bren Professorship of Medical Engineering and Electrical Engineering at the California Institute of Technology. He has published 470 peer-reviewed articles in journals, and has delivered 433 keynotes, plenaries, and invited talks. His Google Scholar h-index and citations have reached 111 and 51,000, respectively.

Biographies of the other 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.
Ruiying Zhang, Lei S. Li, Bin Rao, Haoyang Rong, Min-Yu Sun, Junjie Yao, Ruimin Chen, Qifa Zhou, Steven Mennerick, Baranidharan Raman, and Lihong V. Wang "Multiscale photoacoustic tomography of neural activities with GCaMP calcium indicators," Journal of Biomedical Optics 27(9), 096004 (10 September 2022). https://doi.org/10.1117/1.JBO.27.9.096004
Received: 20 April 2022; Accepted: 24 August 2022; Published: 10 September 2022
Advertisement
Advertisement
KEYWORDS
Brain

Neuroimaging

Photoacoustic tomography

Calcium

Acquisition tracking and pointing

Neurons

Luminescence

Back to Top