Open Access
7 February 2024 Intravital photoacoustic brain stimulation with high-precision
Guangxing Wang, Yuying Zhou, Chunhui Yu, Qiong Yang, Lin Chen, Shuting Ling, Pengyu Chen, Jiwei Xing, Huiling Wu, Qingliang Zhao
Author Affiliations +
Abstract

Significance

Neural regulation at high precision vitally contributes to propelling fundamental understanding in the field of neuroscience and providing innovative clinical treatment options. Recently, photoacoustic brain stimulation has emerged as a cutting-edge method for precise neuromodulation and shows great potential for clinical application.

Aim

The goal of this perspective is to outline the advancements in photoacoustic brain stimulation in recent years. And, we also provide an outlook delineating several prospective paths through which this burgeoning approach may be substantively refined for augmented capability and wider implementations.

Approach

First, the mechanisms of photoacoustic generation as well as the potential mechanisms of photoacoustic brain stimulation are provided and discussed. Then, the state-of-the-art achievements corresponding to this technology are reviewed. Finally, future directions for photoacoustic technology in neuromodulation are provided.

Results

Intensive research endeavors have prompted substantial advancements in photoacoustic brain stimulation, illuminating the unique advantages of this modality for noninvasive and high-precision neuromodulation via a nongenetic way. It is envisaged that further technology optimization and randomized prospective clinical trials will enable a wide acceptance of photoacoustic brain stimulation in clinical practice.

Conclusions

The innovative practice of photoacoustic technology serves as a multifaceted neuromodulation approach, possessing noninvasive, high-accuracy, and nongenetic characteristics. It has a great potential that could considerably enhance not only the fundamental underpinnings of neuroscience research but also its practical implementations in a clinical setting.

1.

Introduction

In the field of neuroscience, neural stimulation has emerged as a vital approach facilitating our comprehension of how the brain functions and the management of neurological conditions. In clinical practice, electrical stimulation, employing implantable devices with metal electrodes, has demonstrated noteworthy efficacy in treating neurological disorders like Alzheimer’s disease, Parkinson’s disease, and epilepsy.13 However, the invasiveness of electrical stimulation based on metal electrodes may potentially lead to unnecessary complications, such as inflammation and bleeding.4 Then, noninvasive neuromodulation strategies, for example, transcranial direct current stimulation and transcranial magnetic stimulation, were developed to circumvent the need for surgical interventions.5,6 Yet, both of them are limited in sophisticated neural circuit manipulation due to the poor spatial (∼several millimeters).7,8 Recently, optogenetics harnessing light to manipulate neural activities via microbial opsins has emerged as a potent method to decipher sophisticated neural circuitries with subcellular spatial resolution and specificity in targeted cell types.911 Nonetheless, the necessity for viral transfection restricts its translation to human subjects.12 Considering the above-mentioned limitations, photothermal neural stimulation as a noninvasive and nongenetic neuromodulation modality has gained significant interest in fundamental neuroscience research and translational studies.1315 Unfortunately, the concomitant thermal toxicity evokes a concern about potential tissue impairment.16 Alternatively, another novel neural manipulation technology, ultrasound neuromodulation has been employed to manipulate the neural activities in the cortex, hippocampus, and thalamus of different species, including mouse,1719 monkey,2022 sheep,23,24 and humans,2527 owing to its noninvasive essence coupled with deep penetration depth.2830 Despite that, during ultrasound brain stimulation, the focus and energy of acoustic wave will be compromised by skull, causing a limited spatial resolution, which fails to meet the demands of single nerve manipulation.31,32 Thus, there is an ongoing quest for novel neuromodulation modalities that aim to realize noninvasive, nongenetic, and high-precision neural manipulation.

In recent years, the photoacoustic technique utilizes pulsed light to generate ultrasound, offering a novel alternative to traditional ultrasound technique, with high penetration depth and spatial precision, leading to rapid advancements in many fundamental and translational studies, particularly in imaging living biological structures across various scales in the life sciences.3337 The photoacoustic technique was also utilized to enhance cell membrane permeability for targeted delivery of normally impermeable molecules, which further expands the capabilities of this technique.38,39 Particularly, in view of the high-precision, and nongenetic merits of photoacoustic technique, Jiang et al. first demonstrated the photoacoustic brain stimulation research via fiber-based photoacoustic emitter.40 Subsequently, various photoacoustic brain stimulation modalities, including photoacoustic film,41 photoacoustic nanotransducer,42 and optically generated focused ultrasound (OFUS),43 have been developed to further enhance the performance of photoacoustic brain stimulation and expand its application scope (Fig. 1). In brief, photoacoustic brain stimulation is a noninvasive and high-precision neuromodulation modality without genetic modification, which has great potential to open up new opportunities for basic neuroscience research and translational studies.

Fig. 1

Photoacoustic brain stimulation methods. (a) Photoacoustic fiber interfaces for neural modulation.40,4446 (b) Photoacoustic film neural stimulation modality based on flexible hydrogel and nanocomposite.41 (c) Photoacoustic neural stimulation based on nanotransducer.42 (d) Neural stimulation based on optically generated focused ultrasound.43 (Adapted with permission from Refs. 38 and 4142.43.)

JBO_29_S1_S11520_f001.png

In this perspective, first, the mechanisms of photoacoustic generation as well as the potential mechanisms of photoacoustic brain stimulation are provided and discussed. Afterward, we summarize the current advancements of photoacoustic brain stimulation in recent years, with an emphasis on their major advances and limitations. Finally, we offer an outlook that highlights a few potential directions for further enhancing the capabilities of this emerging modality, enabling improved performance and wider-ranging applications.

2.

Mechanisms of Photoacoustic Generation

The photoacoustic effect depends on the absorption of pulsed laser light by materials, which results in a transient temperature rise due to non-radiative relaxation processes. This temperature increase leads to a rapid thermal expansion that generates acoustic waves. Two key criteria for generating photoacoustic wave must be met, namely thermal confinement and stress confinement.47 By applying the concept of momentum conservation, a connection can be established between the speed at which thermoelastic expansion occurs and the amplitude of photoacoustic pressure,48 which can be written as

Eq. (1)

P0ρ·vs·V,
where ρ is the mass density (kg/m3), vs is the sound speed (m/s), and V is the speed of thermoelastic volume expansion. In addition, the speed of thermoelastic volume expansion (V) can be expressed as

Eq. (2)

V=ΔVS·τl,
where ΔV is thermoelastic volume expansion (m3), S is the surface area (m2), and τl is laser pulse duration (s). The thermoelastic volume expansion (ΔV) can be defined as

Eq. (3)

ΔV=A·S·F·βρ·Cp,
where A is the light absorption (0<A<1), F is the laser fluence (J/m2), β is volumetric thermal-expansion coefficient (K1), and Cp is the specific heat capacity at constant pressure (J/kg·K). According to the Eqs. (2) and (3), the Eq. (1) can be rewritten as

Eq. (4)

P0=Γ·A·Fl,
where Γ=β·vs2Cp is Grüneisen parameter and l is the characteristic length (m).47 The general consensus is that for achieving high photoacoustic amplitudes, intense light absorption and high thermal expansion are crucial.

On the other hand, to more rigorously estimate the photoacoustic amplitude, the photoacoustic equation has been described as

Eq. (5)

[21vs2·2t2]p(r,t)=ρ·β·2T(r,t)t2,
where p(r,t) represents pressure field; T(r,t) represents temperature field; and ρ·β·2T(r,t)t2 is the acoustic source.49,50 Based on the assumption of negligible heat conduction, the time-dependent temperature field T(r,t) caused by pulsed-laser heating is written as

Eq. (6)

ρCpT(r,t)t=H(r,t),
where H(r,t) is the volumetric nonradiative heat generation caused by light absorption (W/m3). Thus, the Eq. (5) can be rewritten as49

Eq. (7)

[21vs2·2t2]p(r,t)=βCp·H(r,t)t,
where the heating function H(r,t) can be expressed as H(r,t)=I·f(t)·g(r). In this case, I is the peak intensity (W/m2), and f(t) and g(r) are the temporal and spatial heating function, respectively. Therefore, in one-dimensional form, the photoacoustic pressure is approximately described as 49

Eq. (8)

p(t)f*g=f(vsτz)g(z)dz,
where τ represents the retardation time (τ=tz/c); z denotes the distance in the z-direction; * is used to represent the convolution integral; and f and g are defined as temporal heating function and spatial light absorption function, respectively.49 When heat conduction is negligible during optical excitation, the convolution integral becomes an accurate tool for pressure estimation since the spatial heat source aligns with the spatial distribution of light absorption. In this case, f represents a function that varies with the duration of the laser pulse (τl), and g represents a function that varies with the light absorption coefficient (α). In Eq. (8), the photoacoustic generation should be discussed in two situations: thin absorbers and thick absorbers. Assume a light pulse with a Gaussian temporal shape, labeled f(t), is projected onto thin or thick absorbers, each fully absorbing the incoming optical energy. For simplicity, the absorption profile g(z) of each absorber is uniformly distributed (similar to a rectangular function) but differs in penetration depth (1/αthin and 1/αthick for thin and thick absorbers, respectively). Therefore, g(z) can be described as

Eq. (9)

g(z)=g0(H(0)H(z1/α)),
where is the Heaviside function (H(zz0)=0 if z<z0 or 1 if z>z0), and g0 indicates the amplitude of light absorption. In the context of light pulses with uniform fluence F, the light absorption amplitude g0 is considerably larger for the thin absorber compared to the thick absorber, as indicated by F=g0,thin/αthin=g0,thick/αthick Under the assumption of negligible heat conduction during the optical excitation phase, the depth of light absorption and the heat source are identical. According to the convolution integral, the heat source is effectively split into narrow segments, each producing a sound wave that has the same temporal profile as f(t). The combination of these sound waves generates the final photoacoustic wave. In the case of the thin absorber, there are fewer but higher amplitude sound waves, while the thick absorber produces more sound waves, but with lower amplitudes. Regarding the thin absorber, the photoacoustic waves formed have a high amplitude and narrow pulse width (τl+1/vsαthin). In contrast, the photoacoustic waves from the thick absorber present lower amplitudes and wider pulse width (τl+1/vsαthick). Overall, the pulse width of the photoacoustic wave is equivalent to τl+1/vsα. Then, the characteristic length can be calculated as

Eq. (10)

l=vs·τl+1/α.

The photoacoustic pressure amplitude is derived by substituting the Eq. (10) into Eq. (4)

Eq. (11)

P0=Γ·A·Fvs·τl+1/α.

The scenario where vs·τl is significantly greater than 1/α, often found in thin absorbers and expressed as lvs·τl, is known as the long pulse regime. This is because the light pulse duration (τl) is considerably longer than 1/vsα. Conversely, in thick absorbers, when vs·τl is much less than 1/α (implied as l1/α), it’s identified as the short pulse regime. Under these circumstances, the expression for the photoacoustic pressure amplitude is rewritten as47

Eq. (12)

P0={Γ·A·Fvs·τl(vs·τl1/α)Γ·A·F1/α(vs·τl1/α).

3.

Potential Mechanisms of Photoacoustic Brain Stimulation

It is important to note that while photoacoustic technique shows promise as a noninvasive and high-precision neuromodulation modality, its mechanisms are still not fully understood. In essence, photoacoustic brain stimulation utilizes pulsed laser to generate ultrasonic waves for neural stimulation. Several potential mechanisms of photoacoustic brain stimulation are provided for future research directions, including local temperature increase induced by photoacoustic effect,5153 sonoporation,5456 ion channel activation,5759 and intramembrane cavitation,60,61 as shown in Fig. 2.

Fig. 2

Diagram of potential mechanisms for photoacoustic brain stimulation. (a) Local temperature increases on cell membrane induced by photoacoustic effect. (b) Cell membrane pore is induced by photoacoustic effect (sonoporation), which drives ion exchange according to concentration gradients. (c) Ion channels activated by photoacoustic effect contribute to cations influx. (d) Cellular membrane capacitance (ΔCp) changes (intramembrane cavitation) are induced by photoacoustic effect.

JBO_29_S1_S11520_f002.png

The phenomenon of local temperature increase on cell membrane induced by ultrasonic heating was deemed to be the fundamental mechanism with respect to the high-intensity focused ultrasound neuromodulation.62 Instead, investigations involving low-intensity ultrasound have revealed intriguing findings, showing only marginal temperature increases of less than 0.1°C—far below the conventional thermal threshold required for activation (ΔT>5°C).40 In the context of photoacoustic brain stimulation, although the employed pressures and frequencies of it fall within the spectrum of parameters utilized in ultrasound neural stimulation, it is noteworthy that photoacoustic pulses are administered with a duty cycle of 0.36%, where the heat accumulation effect is minimal.40 Briefly, in terms of currently developed photoacoustic brain stimulation modalities, the effect of photoacoustic heating inducing neural activities is negligible.

Another potential mechanism of photoacoustic brain stimulation is sonoporation that essentially leverages the mechanical effects of ultrasound waves to transiently and reversibly break the cellular membrane integrity, which results in ion exchange across the neural membrane and elicits neural activities. Shi et al.38 developed a fiber-based photoacoustic emitter, a novel ultrasound point source that overcomes the acoustic diffraction limitation. Furthermore, this emitter successfully realized the delivery of membrane-impermeable small molecules into living cells via the sonoporation effect, operating under a pressure of 40  kPa and at a low frequency. While its potential has been explored in impermeable molecule delivery, the contribution of sonoporation to altering neural activity through photoacoustic brain stimulation strategy has yet to be fully deciphered. Future investigations utilizing whole-cell electrophysiology may unlock the true contributions of sonoporation in activated neurons during photoacoustic neuromodulation.

Recently, the activation of mechanosensitive ion channels during acoustic neuromodulation has attracted widespread research interest. In Caenorhabditis elegans, Kubanek et al.63 observed MEC-4 (an ion channel required for touch sensation)-dependent currents in vivo during ultrasound modulation. Then, to further test whether mechanical forces can directly induce nerve cell responses, Gaub et al.64 demonstrated that neuronal activity can be modulated by mechanical stimuli through atomic force microscope and calcium imaging technology. Subsequently, Yoo et al.58 scrutinized the activation patterns of diverse mechanosensitive ion channels through ultrasound stimulation and calcium fluorescent imaging. Consequently, they discerned the pivotal involvement of three distinct ion channels: namely, TRPP2, TRPC1, and Piezo1. Nevertheless, the electrophysiological investigations at the single neuron level are limited due to the inapplicability of whole-cell recording with ultrasound stimulation. Fortunately, a tapered fiber photoacoustic emitter developed by Shi et al. is capable of stimulating single neuron or only subcellular structures, which finally makes the integration of photoacoustic stimulation with patch-clamp recording on single neuron feasible. Thus, the detailed ion channel dynamics involved in mechanical stimulation by photoacoustic neuromodulation can be further unveiled in the future.

Another prevailing explanation of how ultrasound activates neurons is intramembrane cavitation, which disturbs the structure of neural membrane and induces capacitive currents. Krasovitski et al.61 constructed a “bilayer sonophore” model to study how the mechanical energy of ultrasound is absorbed by the cellular membrane and induces intramembrane cavitation. The results demonstrated that the phenomenon of intramembrane cavitation was observed under the condition of continuous wave ultrasound at the frequency of 1 MHz. Then on the basis of this study, Plaksin et al.60 further verified that the ultrasound-induced intramembrane cavitation is capable of leading to neuron excitation via the effect of currents induced by changes of membrane capacitance. However, the cultured primary cortical neurons were successfully evoked by single-cycle and broad bandwidth photoacoustic wave generated by OFUS.43 Thus, the intramembrane cavitation mechanism may not be applicable to photoacoustic neuromodulation.

4.

Photoacoustic Brain Stimulation Modalities

4.1.

Fiber-Based Photoacoustic Emitters

Lately, fiber-based photoacoustic emitters were developed as a miniature ultrasound source for all-optical ultrasound imaging and surgical guidance.65,66 Beyond these applications, Jiang et al.40 first demonstrated the innovative application of fiber-based photoacoustic emitter modality in neuromodulation at submillimeter spatial precision [Figs. 3(a) and 3(b)]. The fiber-based photoacoustic emitter termed as fiber-optoacoustic converter (FOC) fabricated by coating a fiber tip with a light diffusion layer (ZnO/epoxy mixture) and an absorption layer (graphite/epoxy mixture) in this study has a diameter of 600  μm and can activate neurons within a radius of 500  μm, providing superior spatial resolution compared to conventional ultrasound neuromodulation. Calcium transients were observed in response to laser pulse trains delivered by the FOC, and no morphological changes were detected in the stimulated neurons [Figs. 3(c) and 3(d)]. Additionally, the FOC was able to evoke motor responses with high spatial precision in the motor cortex. No response was observed in the contralateral A1, indicating that the auditory pathway was not involved in the neural activation. In brief, the FOC has shown promising potential for high-precision neural stimulation without the need for genetic modification.

Fig. 3

Fiber-based photoacoustic emitters. (a) The diagram of photoacoustic neuromodulation through a FOE. Inset is the enlarged FOE tip. (b) Schematic of acoustic wave generation. (c) Photoacoustic neuromodulation induced calcium transients in cultured primary neurons loaded with OGD-1. (d) Calcium trace of a neuron undergone repeated FOE stimulation. Green arrow: stimulation onset. (Adapted with permission from Ref. 40.) (e) Multiwall CNT/PDMS mixture as coating material casted on a metal mesh followed by a punch-through method to coat the tapered fiber. (f) Detected pressure plotted as a function of the distance. (g) TFOE-induced stimulation of GCaMP6f expressing single neuron. Scale bar: 50  μm. (h) TFOE selectively stimulation of axon (red) and dendrites (yellow and green) of a multipolar neuron. Scale bar: 50  μm. (Adapted with permission from Ref. 44.) (i) Key steps of CSFOE fabrication. Scale bars: 200 mm. (j) Diagram of dual site stimulation using two CSFOEs with a fiber splitter. (k) Map of the max ΔF/F0 image of two sites of neurons stimulated by two CSFOE. (Adapted with permission from Ref. 45.) (l) Diagram of mFOE for bidirectional communication with neurons. (m) Illustration of the mFOE enabled bidirectional neural communication using laser signal as input and electrical signal as readout. (p) mFOE was implanted into hippocampus of a wild type C57BL/6J mouse. Simultaneous optoacoustic stimulation and electrophysiological recording performed at 3 days (n), 7 days (o), 2 weeks (q), and 1 month (r) after implantation. (Adapted with permission from Ref. 46.)

JBO_29_S1_S11520_f003.png

To further improve the spatial precision of photoacoustic stimulation, Shi et al.44 proposed a further miniaturized fiber-based photoacoustic emitter termed as tapered fiber optoacoustic emitter (TFOE), which is capable of manipulating a single neuron with an unprecedented high spatial precision [Figs. 3(e)3(h)]. The researchers fabricated the TFOE with a diameter of 20 μm at the tip [Fig. 3(e)], in which the absorption/thermal expansion layer using carbon nanotubes (CNTs) embedded in polydimethylsiloxane (PDMS) was optimized to improve photoacoustic conversion efficiency. The spatial resolution of the acoustic field generated by TFOE was found to be 39.6  μm [Fig. 3(f)], satisfying single-cell manipulation. Importantly, TFOE stimulation successfully targeted subcellular structures, such as axons and dendrites, within neurons [Figs. 3(g) and 3(h)]. Thus, these results revealed that TFOE has neurostimulation capabilities with high accuracy and reliability, providing new possibilities for neuroscience studies at the level of individual neurons and potential clinical applications without genetic modifications.

Generally, multiple functional regions of the brain are always involved in sophisticated brain functions. Thus, a high-precision and multi-site stimulation tool is needed. For achieving this goal, based on their previous study,40 Chen et al.45 developed a new fiber-based photoacoustic emitter named as candle soot-based fiber optoacoustic emitters (CSFOE) with high photoacoustic conversion efficiency [Figs. 3(i)3(k)]. The CSFOE was fabricated by coating the tip of a polished multimode optical fiber with candle soot synthesized from a paraffin wax candle, followed by coating with PDMS using a nanoinjector [Fig. 3(i)]. Besides, the pressure of the generated acoustic signal by CSFOE reached approximately 10 MPa, which is 9.6 times larger compared with that generated by FOC. Based on this advantage, CSFOE successfully realized dual-site neuron stimulation with an average maximum fluorescence change of over 10% in GCaMP6f-labeled neuron cultures [Figs. 3(j) and 3(k)]. Therefore, the CSFOE had superior spatial resolution and high-pressure conversion efficiency, making it suitable for modulating complex animal behavior by controlling multiple target sites in the brain circuitry.

Bidirectional communication with neural circuits in the brain is crucial for fundamental studies and clinical treatments of neurological diseases. However, existing methods such as electrical stimulation and optogenetics have limitations in terms of interference with electrical recording, low efficiency in viral transfection, and safety concerns, respectively. To overcome the limitations of existing methods, Zheng et al.46 developed a multifunctional fiber-based optoacoustic emitter (mFOE) that combines photoacoustic neuromodulation and electrical recording [Fig. 3(l)], which is orthogonal to electrical recording and does not require viral transfection, making it a promising candidate for bidirectional brain interfaces [Figs. 3(l)3(r)].

4.2.

Photoacoustic Film

Utilizing biocompatible scaffolds as neural interfaces is crucial for the functional repair of nerve injuries and rehabilitation of neurodegenerative diseases. Furthermore, neural stimulation has been found to promote neural regeneration.67 Therefore, various factors such as mechanical68 and chemical stimuli69 were involved in functionalizing nerve scaffolds, in which electrical stimulation70,71 is the most widely applied technique. However, the delivery of electrical stimulus to conductive scaffolds remains challenging, and current solutions have limitations in terms of spatial resolution and the risk of infection. Aiming to overcome these limitations, Zheng et al.41 proposed a novel photoacoustic neural stimulation modality, a flexible and biocompatible photoacoustic film, for promoting neural regeneration (Fig. 4). This photoacoustic film was fabricated by embedding functioned CNTs, efficient photoacoustic agent, into silk fibroin solution, an FDA-approved biocompatible material, and casting the mixture [Fig. 4(a)]. The viability of cortical neurons cultured on photoacoustic film was evaluated using the MTS assay, and no significant difference in cell viability was observed compared to the control group [Fig. 4(b)]. Neurons cultured on photoacoustic film showed increased fluorescence intensity after photoacoustic stimulation [Figs. 4(c) and 4(d)]. Furthermore, by promoting the secretion of brain-derived neurotrophic factors (BDNF), this innovative photoacoustic stimulation modality has been proven capable of facilitating neural regeneration [Figs. 4(e) and 4(f)]. Compared with other neuromodulation modalities, such as electrical stimulation and optogenetics, this innovative photoacoustic film eliminates the need for cumbersome wire connections and genetic modifications, making it a convenient and versatile option for researchers and clinicians. Moreover, photoacoustic film is complementary to other photoacoustic brain stimulation modalities like fiber-based photoacoustic emitters, further expanding its potential applications in the field of neural stimulation.

Fig. 4

Flexible and biocompatible photoacoustic film for neural stimulation and regeneration. (a) Diagram of the fabrication process of photoacoustic film. (b) Biological safety of silk film (as control), CNT/silk film, and silk film with freeform CNT. Calcium images of rat cortical neurons (d) before and (c) after photoacoustic neural stimulation. (e) Average neurite coverage area for dorsal root ganglion cells in four groups. (f) Average concentrations of BDNF of photoacoustic stimulated and unstimulated dorsal root ganglion cells. (Adapted with permission from Ref. 41.)

JBO_29_S1_S11520_f004.png

4.3.

Photoacoustic Nanotransducer

Recently, there has been a remarkable surge in the development of nanoparticle-assisted neuromodulation techniques.7274 Particularly, semiconducting polymer nanoparticles exhibit unique advantages, such as the remarkable ability to absorb near-infrared light, ensuring optimal biocompatibility, and allowing for controlled biodegradation.75 Building upon this, Jiang et al.42 innovatively developed a photoacoustic brain stimulation nanocomposite platform, termed as photoacoustic nanotransducer, which was created using bis-isoindigo-based polymer (BTII) and modified with poly(styrene)-b-poly(acrylic acid) (PS-b-PAA) forming water-soluble nanoparticles (50  nm) via nanoprecipitation (Fig. 5). In vitro experiments were successfully conducted to activate the primary neurons using nanosecond laser pulses at 1030 nm [Figs. 5(b) and 5(c)], which have been verified that the high temporal resolution (∼ms) and single-cell spatial resolution can be realized. Besides, the stimulation specificity of photoacoustic nanotransducer was further achieved by conjugating the mechanosensitive ion channel TRPV4 antibody with photoacoustic nanotransducer to target the mechanosensitive TRPV4 channels on the neuronal membrane. Furthermore, in vivo experiments were performed by injecting photoacoustic nanotransducer directly into the brain to activate the motor cortex and subsequent motor responses were observed via electromyography (EMG) recordings. Briefly, harnessing NIR-II excitation, the nongenetic neuromodulation modality is capable of achieving deep tissue penetration and targeting cellular specificity stimulation.

Fig. 5

Nanotransducer-mediated photoacoustic brain stimulation. (a) Diagram of photoacoustic nanotransducer induced neural stimulation (left) and the PAN generating photoacoustic signal generated by illuminating photoacoustic nanotransducer with nanosecond laser pulses (right). (b) Calcium images of neurons transfected by GCaMP6f and cultured with photoacoustic nanotransducer for 15 min. White circle: illumination position. (c) Colormaps of fluorescence changes of neurons stimulated by photoacoustic nanotransducer. (d) Diagram of in vivo neural stimulation by injected photoacoustic nanotransducer coupled with electrophysiology measurement. (e) Electrophysiology curves recorded at the brain region without photoacoustic nanotransducer as the control (blue) and photoacoustic nanotransducer treated region (red). Blue arrow: stimulation onset. (Adapted with permission from Ref. 42.)

JBO_29_S1_S11520_f005.png

4.4.

Optically Generated Focused Ultrasound by Curved Soft Optoacoustic Pad

While previously developed fiber-based photoacoustic emitters neuromodulation modalities are capable of achieving non-genetic, high-precision neural stimulation, a surgical implantation procedure is usually required because the fiber-based photoacoustic emitters utilize near-field ultrasound for localized neural stimulation. To overcome this limitation, Li et al.43 proposed a novel photoacoustic brain stimulation modality called OFUS for noninvasive brain stimulation with ultrahigh precision (Fig. 6). In this study, the OFUS was generated by a curved soft optoacoustic pad (SOAP), fabricated by candle soot layered with PDMS, upon a pulsed laser illumination [Fig. 6(a)]. OFUS produced an ultrahigh spatial resolution of approximately 83  μm, significantly better than transcranial-focused ultrasound (tFUS). Upon the illumination of a pulsed laser, OFUS successfully evoked neuron excitation observed by calcium imaging in vitro, which demonstrated that OFUS has the ability to evoke responses in neurons and achieve localized stimulation [Fig. 6(b)]. Additionally, the transcranial stimulation capability of OFUS generated by SOAP was also investigated in vivo. The stimulation results were evaluated by both immunofluorescence imaging (c-Fos) and electrophysiology recording [Figs. 6(e)6(h)]. Significantly, the c-Fos signal remained localized exclusively within the designated target site, encompassing an approximate area of 200  μm in diameter. This outcome effectively showcases a superior spatial resolution compared to the conventional tFUS stimulation (15  mm).76 Collectively, OFUS provides exceptionally precise non-invasive methodologies for delving into neurological investigations within the sub-regions of the brain, which has great potential to be a crucial technology for advancing both neuroscience research and clinical interventions.

Fig. 6

OFUS for neuromodulation. (a) The diagram of OFUS design. Numerical aperture and lateral resolutions (b) and numerical aperture and axial resolutions (c). Orange area: the NA range of conventional ultrasound transducers. (d) Calcium images of neuron activities before and after OFUS stimulation. Scale bar: 50  μm. (e) The diagram of OFUS in vivo. (f) Statistic analysis of the percentage of c-Fos positive neurons after OFUS stimulation. (g) EMG recordings of 2 s OFUS stimulation at the somatosensory cortex. Orange box: laser on. (h) EMG signals after the band-pass filter and full-wave rectifier and envelope. (Adapted with permission from Ref. 43.)

JBO_29_S1_S11520_f006.png

5.

Outlook and Future Directions

As previously deliberated, photoacoustic brain stimulation emerging as a novel and multifaceted modality has posed great potential to propel the domain of acoustic neuromodulation forward, spanning not only fundamental scientific exploration but also intricate clinical utilizations. While utilizing the emerging photoacoustic neuromodulation approach represents a promising advancement in addressing prior challenges, its current stage reveals limitations that underline the need for ongoing improvements in the field. Aiming to achieve high-precision and non-genetic neuromodulation, fiber-based photoacoustic emitters photoacoustic stimulation modalities including FOC, TFOE, CSFOE, and mFOE were developed and characterized by single neuron stimulation, sub-cellular stimulation, multi-site stimulation, and bidirectional communications, respectively. Unfortunately, these modalities require surgical implantation into the target area and are not suitable for transcranial application. Besides, the future development direction of fiber-optic stimulators is heading towards multifunctionality. Thus, in future work, one can combine pharmacological intervention channel with this fiber-based photoacoustic stimulation modality to construct a multifunctional fiber-based photoacoustic emitter for achieving more sophisticated brain research. In addition, to tackle the problem of different form of neural interface requirements in neuromodulation areas such as the brain cortex, retina, and peripheral nerve, a flexible and biocompatible scaffold with photoacoustic properties was developed. The photoacoustic film, when excited with a 1030 nm pulsed laser, produced a broadband photoacoustic wave, initiating a calcium influx in neurons and facilitating in the proliferation of neurites. Nevertheless, the process of implanting photoacoustic film is still invasive and might not be practical for larger-volume condition. Therefore, using injectable materials could be a potential answer to address this clinical challenge.77 Moreover, photoacoustic nanotransducer, a new nongenetic nanoparticle-assisted neuromodulation platform, was created to achieve neural activation with enhanced specificity by conjugating with TRPV4 antibody. The functionality of photoacoustic nanotransducer for neuromodulation in a living organism was confirmed by its direct administration into the motor cortex of a mouse and triggering it with an NIR II laser light. However, the delivery process of photoacoustic nanotransducer was invasive. And, the penetration depth of NIR II laser light utilized in this study was limited in deep cerebral nuclei stimulation. New photoacoustic nanotransducer, conjugating with specific neural cell membrane channel protein antibody, with capabilities of penetrating blood brain barrier, or noninvasive blood brain barrier opening delivery method, and advanced optical wavefront shaping method for deep tissue focus should be developed to realize cell-specific, nongenetic and noninvasive photoacoustic nanotransducer neuromodulation in future. Furthermore, as an exogenous agent, the potential toxicity of photoacoustic nanotransducer introduced into animals and eventually in patients should be further confirmed in subsequent research. On the other hand, to realize high-precision photoacoustic neuromodulation without any surgical implantation, the OFUS generated by SOAP was created, which can perform high-precision transcranial neuromodulation compared with traditional ultrasound stimulation. However, the spatial resolution of SOAP is inferior to that of fiber-based photoacoustic emitters. That is caused by the distortion of thick skulls. Therefore, in future work, acoustic wavefront engineering should be done to compensate for this aberration. In summary, different photoacoustic neuromodulation platforms possess distinct characteristics and advantages. Fiber-based photoacoustic emitter modalities, such as TFOE and mFOE, exhibit strengths in the field of single cell/subcellular stimulation and bidirectional communication (i.e., carrying out both neural stimulation and simultaneous electrical recording of neural responses). With its outstanding biocompatibility and flexibility, the photoacoustic film is adept at forming conformal attachments to tissues with varying shapes. Moreover, the photoacoustic nanotransducer, integrated with an antibody coupling strategy, opens up possibilities for precisely targeted stimulation of specific cell type. In addition, OFUS is capable of performing noninvasive and ultrahigh precision (below 0.1 mm) neuromodulation without surgical implantation, proving to be an essential technology in the fields of neuroscience research and clinical therapy. The pros and cons of four different photoacoustic neuromodulation platforms are shown in Table 1.

Table 1

Pros and cons of four different photoacoustic neuromodulation platforms.

PlatformProsCons
Fiber-based photoacoustic emittersSingle cell/subcellular precision, and bidirectional communicationInvasive
Photoacoustic filmBiocompatible and flexibleImplantation process remains invasive
Photoacoustic nanotransducerSpecific targetingInvasive delivery, limited depth, and biosafety should be further confirmed
OFUSNon-invasive transcranial neural stimulation with ultrahigh precision (below 0.1 mm)Skull distortion
OFUS: optically generated focused ultrasound.

Ensuring the brain’s safety with photoacoustic wave is vital for the feasibility of photoacoustic neuromodulation as an effective brain stimulation technique. Mechanical and thermal effects are the main safety concerns during photoacoustic stimulation. First, the mechanical index (MI), a non-dimensional measure widely employed in the field of ultrasound, serves to estimate the potential for mechanical damage caused by ultrasound. It is represented as52

Eq. (13)

MI=PnegativeFc,
where Pnegative is peak negative pressure of the ultrasound wave (MPa) and Fc is the center frequency of the ultrasound pulse (MHz). In the case of mFOE and OFUS, the MI of generated photoacoustic wave is 0.2 and 0.5, respectively, which is below 1.9, the recommended safety limit set by the FDA guidelines.46,78 Moreover, the peak negative pressure of FOC, TFOE, CSFOE, photoacoustic film, photoacoustic nanotransducer was estimated below the threshold of bubble cloud generation in soft tissue (25 to 30 MPa).79

For thermal safety, in FOC, TFOE, CSFOE, mFOE, photoacoustic film, OFUS platform, the maximum temperature increase does not exceed 1.6°C. Such temperature increase is well below the previously reported threshold for thermal induced neural damage.80 However, in photoacoustic nanotransducer platform, a rapid temperature surge was simulated, peaking at 8.4°C at the photoacoustic nanotransducer surface and attaining 5.0°C 10 nm away from the surface, which may introduce thermotoxicity during chronic in vivo stimulation under body temperature. Further thermal safety research is needed in future. These initial studies are promising, but there is a need for more extensive studies to fully determine the safety and efficacy of photoacoustic neuromodulation for immediate applications in the brain, and also to set safety guidelines for upcoming chronic applications.

Different frequencies used in ultrasound stimulation produce distinct results due to its frequency-specific nature. For example, prior research81 has demonstrated variability in the effectiveness of neuron spiking induction by ultrasound neurostimulation in mice, across a frequency spectrum of 0.3 to 2.9 MHz. Higher frequencies in this range require increased spatial peak intensities to maintain the same effectiveness as lower frequencies. In addition, neural inhibition effects were induced using high frequency ultrasound operating at 30 MHz.82 Nevertheless, due to the broad bandwidth nature of the photoacoustic wave, currently, it is limited to isolate frequency-specific responses during photoacoustic neuromodulation. So far, no studies have shown that photoacoustic neuromodulation technology can hyperpolarize neurons to inhibit their activity. Notably, brain network connectivity and activities are highly complex, and stimulating or inhibiting these neuronal activities is vital to comprehend their functions. Limitation of isolating frequency-specific responses and absent the capacity to suppress the activity of neurons impeded the study of complex neural activities by photoacoustic neuromodulation technology. On the other hand, the advantage of photoacoustic broadband stimulation relative to specific frequency ultrasound stimulation is that photoacoustic neuromodulation can stimulate neuronal activity using acoustic waves with shorter pulse duration (1  μs) than those used in ultrasound stimulation. Therefore, the relative advantages and disadvantages of broadband stimulation in photoacoustic stimulation should be carefully considered depending on the specific application and research goals. Additionally, to date, there is no clear understanding about the mechanism of photoacoustic stimulation. Fortunately, TFOE is metal-free and compatible with patch clamp recording technology. Thus, future investigations utilizing TFOE, electrophysiological recording, genetic and pharmacological intervention could provide insight into the molecular mechanism of photoacoustic stimulation.

Disclosures

The authors declare no conflicts of interest.

Code and Data Availability

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

Acknowledgments

This work was supported by grants from the Guangdong Basic and Applied Basic Research Foundation (Grant No. 2021A1515011654); Fundamental Research Funds for the Central Universities of China (Grant No. 20720210117); Fujian Province Science and Technology Plan Guiding Project (Grant No. 2022Y0002); Science and Technology Projects Innovation Laboratory for Sciences and Technologies of Energy Materials of Fujian Province (IKKEM) (Grant No. RD2022050901); Fundamental Research Funds for the Central Universities (Grant No. 20720232020); and XMU Undergraduate Innovation and Entrepreneurship Training Programs (Grant Nos. 2020Y0799, S202010384334, 2021X1119, 2021Y111, S202110384391, S202210384404, and 202210384051).

References

1. 

C. Yang et al., “Synergistic effect of electric stimulation and mesenchymal stem cells against Parkinson’s disease,” Aging, 12 (16), 16062 https://doi.org/10.18632/aging.103477 AGNYDE 0160-2721 (2020). Google Scholar

2. 

B. J. Stieve et al., “Optimization of closed-loop electrical stimulation enables robust cerebellar-directed seizure control,” Brain, 146 (1), 91 –108 https://doi.org/10.1093/brain/awac051 BRAIAK 0006-8950 (2023). Google Scholar

3. 

A. Majdi et al., “A systematic review and meta-analysis of transcranial direct-current stimulation effects on cognitive function in patients with Alzheimer’s disease,” Mol. Psychiatry, 27 (4), 2000 –2009 https://doi.org/10.1038/s41380-022-01444-7 (2022). Google Scholar

4. 

G. Taccola et al., “Complications of epidural spinal stimulation: lessons from the past and alternatives for the future,” Spinal Cord, 58 (10), 1049 –1059 https://doi.org/10.1038/s41393-020-0505-8 (2020). Google Scholar

5. 

P. Sudbrack-Oliveira et al., “Transcranial direct current stimulation (tDCS) in the management of epilepsy: a systematic review,” Seizure, 86 85 –95 https://doi.org/10.1016/j.seizure.2021.01.020 SEIZE7 1059-1311 (2021). Google Scholar

6. 

V. Di Lazzaro et al., “Diagnostic contribution and therapeutic perspectives of transcranial magnetic stimulation in dementia,” Clin. Neurophysiol., 132 (10), 2568 –2607 https://doi.org/10.1016/j.clinph.2021.05.035 CNEUFU 1388-2457 (2021). Google Scholar

7. 

J. Giordano et al., “Mechanisms and effects of transcranial direct current stimulation,” Dose-Response, 15 (1), 1559325816685467 https://doi.org/10.1177/1559325816685467 (2017). Google Scholar

8. 

T. Yoshikawa et al., “Temporal and spatial profiles of evoked activity induced by magnetic stimulation using millimeter-sized coils in the mouse auditory cortex in vivo,” Brain Res., 1796 148092 https://doi.org/10.1016/j.brainres.2022.148092 BRREAP 0006-8993 (2022). Google Scholar

9. 

V. Emiliani et al., “Optogenetics for light control of biological systems,” Nat. Rev. Methods Prim., 2 (1), 55 https://doi.org/10.1038/s43586-022-00136-4 (2022). Google Scholar

10. 

M. Ledri et al., “Optogenetics for controlling seizure circuits for translational approaches,” Neurobiol. Dis., 184 106234 https://doi.org/10.1016/j.nbd.2023.106234 NUDIEM 0969-9961 (2023). Google Scholar

11. 

S. Kim et al., “Whole-brain mapping of effective connectivity by fMRI with cortex-wide patterned optogenetics,” Neuron, 111 (11), 1732 –1747.e6 https://doi.org/10.1016/j.neuron.2023.03.002 NERNET 0896-6273 (2023). Google Scholar

12. 

M. White, M. Mackay and R. G. Whittaker, “Taking optogenetics into the human brain: opportunities and challenges in clinical trial design,” Open Access J. Clin. Trials, 12 33 –41 https://doi.org/10.2147/OAJCT.S259702 (2020). Google Scholar

13. 

Y. An and Y. Nam, “Closed-loop control of neural spike rate of cultured neurons using a thermoplasmonics-based photothermal neural stimulation,” J. Neural Eng., 18 (6), 066002 https://doi.org/10.1088/1741-2552/ac3265 1741-2560 (2021). Google Scholar

14. 

X. Wu et al., “Tether-free photothermal deep-brain stimulation in freely behaving mice via wide-field illumination in the near-infrared-II window,” Nat. Biomed. Eng., 6 (6), 754 –770 https://doi.org/10.1038/s41551-022-00862-w (2022). Google Scholar

15. 

S. Yoo, J.-H. Park and Y. Nam, “Single-cell photothermal neuromodulation for functional mapping of neural networks,” ACS Nano, 13 (1), 544 –551 https://doi.org/10.1021/acsnano.8b07277 ANCAC3 1936-0851 (2018). Google Scholar

16. 

M. M. Chernov, G. Chen and A. W. Roe, “Histological assessment of thermal damage in the brain following infrared neural stimulation,” Brain Stimul., 7 (3), 476 –482 https://doi.org/10.1016/j.brs.2014.01.006 (2014). Google Scholar

17. 

X. Wang et al., “Neuromodulation effects of ultrasound stimulation under different parameters on mouse motor cortex,” IEEE Trans. Biomed. Eng., 67 (1), 291 –297 https://doi.org/10.1109/TBME.2019.2912840 IEBEAX 0018-9294 (2019). Google Scholar

18. 

S. Dong et al., “Modulation effect of mouse hippocampal neural oscillations by closed-loop transcranial ultrasound stimulation,” J. Neural Eng., 19 (6), 066038 https://doi.org/10.1088/1741-2552/aca799 1741-2560 (2022). Google Scholar

19. 

X. Wang et al., “Ultrasonic thalamic stimulation modulates neural activity of thalamus and motor cortex in the mouse,” J. Neural Eng., 18 (6), 066037 https://doi.org/10.1088/1741-2552/ac409f 1741-2560 (2021). Google Scholar

20. 

P.-F. Yang et al., “Neuromodulation of sensory networks in monkey brain by focused ultrasound with MRI guidance and detection,” Sci. Rep., 8 (1), 7993 https://doi.org/10.1038/s41598-018-26287-7 SRCEC3 2045-2322 (2018). Google Scholar

21. 

D. Folloni et al., “Manipulation of subcortical and deep cortical activity in the primate brain using transcranial focused ultrasound stimulation,” Neuron, 101 (6), 1109 –1116.e5 https://doi.org/10.1016/j.neuron.2019.01.019 NERNET 0896-6273 (2019). Google Scholar

22. 

D. Folloni et al., “Ultrasound modulation of macaque prefrontal cortex selectively alters credit assignment–related activity and behavior,” Sci. Adv., 7 (51), eabg7700 https://doi.org/10.1126/sciadv.abg7700 STAMCV 1468-6996 (2021). Google Scholar

23. 

P. Gaur et al., “Histologic safety of transcranial focused ultrasound neuromodulation and magnetic resonance acoustic radiation force imaging in rhesus macaques and sheep,” Brain Stimul., 13 (3), 804 –814 https://doi.org/10.1016/j.brs.2020.02.017 (2020). Google Scholar

24. 

W. Lee et al., “Image-guided focused ultrasound-mediated regional brain stimulation in sheep,” Ultrasound Med. Biol., 42 (2), 459 –470 https://doi.org/10.1016/j.ultrasmedbio.2015.10.001 USMBA3 0301-5629 (2016). Google Scholar

25. 

A. Fomenko et al., “Low-intensity ultrasound neuromodulation: an overview of mechanisms and emerging human applications,” Brain Stimul., 11 (6), 1209 –1217 https://doi.org/10.1016/j.brs.2018.08.013 (2018). Google Scholar

26. 

J. M. Stern et al., “Safety of focused ultrasound neuromodulation in humans with temporal lobe epilepsy,” Brain Stimul., 14 (4), 1022 –1031 https://doi.org/10.1016/j.brs.2021.06.003 (2021). Google Scholar

27. 

Y. Zhang et al., “Transcranial ultrasound stimulation of the human motor cortex,” iScience, 24 (12), 103429 https://doi.org/10.1016/j.isci.2021.103429 (2021). Google Scholar

28. 

W. Lee et al., “Safety review and perspectives of transcranial focused ultrasound brain stimulation,” Brain Neurorehabil., 14 (1), 1 –16 https://doi.org/10.12786/bn.2021.14.e4 (2021). Google Scholar

29. 

P. Bowary and B. D. Greenberg, “Noninvasive focused ultrasound for neuromodulation: a review,” Psychiatr. Clin., 41 (3), 505 –514 https://doi.org/10.1016/j.psc.2018.04.010 PSCLBV 0033-264X (2018). Google Scholar

30. 

H. Baek, K. J. Pahk and H. Kim, “A review of low-intensity focused ultrasound for neuromodulation,” Biomed. Eng. Lett., 7 (2), 135 –142 https://doi.org/10.1007/s13534-016-0007-y (2017). Google Scholar

31. 

L. Di Biase, E. Falato and V. Di Lazzaro, “Transcranial focused ultrasound (tFUS) and transcranial unfocused ultrasound (tUS) neuromodulation: from theoretical principles to stimulation practices,” Front. Neurol., 10 549 https://doi.org/10.3389/fneur.2019.00549 (2019). Google Scholar

32. 

P.-C. Tsai, H. S. Gougheri and M. Kiani, “Skull impact on the ultrasound beam profile of transcranial focused ultrasound stimulation,” in 41st Annu. Int. Conf. of the IEEE Eng. in Med. and Biol. Soc. (EMBC), 5188 –5191 (2019). https://doi.org/10.1109/EMBC.2019.8857269 Google Scholar

33. 

I. Steinberg et al., “Photoacoustic clinical imaging,” Photoacoustics, 14 77 –98 https://doi.org/10.1016/j.pacs.2019.05.001 (2019). Google Scholar

34. 

D. Das et al., “Another decade of photoacoustic imaging,” Phys. Med. Biol., 66 (5), 05TR01 https://doi.org/10.1088/1361-6560/abd669 PHMBA7 0031-9155 (2021). Google Scholar

35. 

J. Weber, P. C. Beard and S. E. Bohndiek, “Contrast agents for molecular photoacoustic imaging,” Nat. Methods, 13 (8), 639 –650 https://doi.org/10.1038/nmeth.3929 1548-7091 (2016). Google Scholar

36. 

L. V. Wang and S. Hu, “Photoacoustic tomography: in vivo imaging from organelles to organs,” Science, 335 (6075), 1458 –1462 https://doi.org/10.1126/science.1216210 SCIEAS 0036-8075 (2012). Google Scholar

37. 

Z. Hosseinaee et al., “Towards non-contact photoacoustic imaging,” Photoacoustics, 20 100207 https://doi.org/10.1016/j.pacs.2020.100207 (2020). Google Scholar

38. 

L. Shi et al., “A fiber optoacoustic emitter with controlled ultrasound frequency for cell membrane sonoporation at submillimeter spatial resolution,” Photoacoustics, 20 100208 https://doi.org/10.1016/j.pacs.2020.100208 (2020). Google Scholar

39. 

A. Sengupta et al., “Poloxamer surfactant preserves cell viability during photoacoustic delivery of molecules into cells,” Biotechnol. Bioeng., 112 (2), 405 –415 https://doi.org/10.1002/bit.25363 BIBIAU 0006-3592 (2015). Google Scholar

40. 

Y. Jiang et al., “Optoacoustic brain stimulation at submillimeter spatial precision,” Nat. Commun., 11 (1), 881 https://doi.org/10.1038/s41467-020-14706-1 NCAOBW 2041-1723 (2020). Google Scholar

41. 

N. Zheng et al., “Photoacoustic carbon nanotubes embedded silk scaffolds for neural stimulation and regeneration,” ACS Nano, 16 (2), 2292 –2305 https://doi.org/10.1021/acsnano.1c08491 ANCAC3 1936-0851 (2022). Google Scholar

42. 

Y. Jiang et al., “Neural stimulation in vitro and in vivo by photoacoustic nanotransducers,” Matter, 4 (2), 654 –674 https://doi.org/10.1016/j.matt.2020.11.019 (2021). Google Scholar

43. 

Y. Li et al., “Optically-generated focused ultrasound for noninvasive brain stimulation with ultrahigh precision,” Light Sci. Appl., 11 (1), 321 https://doi.org/10.1038/s41377-022-01004-2 (2022). Google Scholar

44. 

L. Shi et al., “Non-genetic photoacoustic stimulation of single neurons by a tapered fiber optoacoustic emitter,” Light Sci. Appl., 10 (1), 143 https://doi.org/10.1038/s41377-021-00580-z (2021). Google Scholar

45. 

G. Chen et al., “High-precision neural stimulation by a highly efficient candle soot fiber optoacoustic emitter,” Front. Neurosci., 16 1005810 1662-453X10.3389/fnins.2022.1005810 (2022). Google Scholar

46. 

N. Zheng et al., “Multifunctional fiber-based optoacoustic emitter as a bidirectional brain interface,” Adv. Healthc. Mater., 12 (25), 2300430 https://doi.org/10.1002/adhm.202300430 (2023). Google Scholar

47. 

T. Lee et al., “Efficient photoacoustic conversion in optical nanomaterials and composites,” Adv. Opt. Mater., 6 (24), 1800491 https://doi.org/10.1002/adom.201800491 2195-1071 (2018). Google Scholar

48. 

D. Kim, M. Ye and C. Grigoropoulos, “Pulsed laser-induced ablation of absorbing liquids and acoustic-transient generation,” Appl. Phys. A, 67 169 –181 https://doi.org/10.1007/s003390050756 (1998). Google Scholar

49. 

G. Diebold, T. Sun and M. Khan, “Photoacoustic monopole radiation in one, two, and three dimensions,” Phys. Rev. Lett., 67 (24), 3384 https://doi.org/10.1103/PhysRevLett.67.3384 PRLTAO 0031-9007 (1991). Google Scholar

50. 

M. Xu and L. V. Wang, “Photoacoustic imaging in biomedicine,” Rev. Sci. Instrum., 77 (4), 041101 https://doi.org/10.1063/1.2195024 RSINAK 0034-6748 (2006). Google Scholar

51. 

H. A. Kamimura et al., “Ultrasound neuromodulation: mechanisms and the potential of multimodal stimulation for neuronal function assessment,” Front. Phys., 8 150 https://doi.org/10.3389/fphy.2020.00150 (2020). Google Scholar

52. 

J. Blackmore et al., “Ultrasound neuromodulation: a review of results, mechanisms and safety,” Ultrasound Med. Biol., 45 (7), 1509 –1536 https://doi.org/10.1016/j.ultrasmedbio.2018.12.015 USMBA3 0301-5629 (2019). Google Scholar

53. 

C. Constans et al., “Potential impact of thermal effects during ultrasonic neurostimulation: retrospective numerical estimation of temperature elevation in seven rodent setups,” Phys. Med. Biol., 63 (2), 025003 https://doi.org/10.1088/1361-6560/aaa15c PHMBA7 0031-9155 (2018). Google Scholar

54. 

B. Helfield et al., “Biophysical insight into mechanisms of sonoporation,” Proc. Natl. Acad. Sci., 113 (36), 9983 –9988 https://doi.org/10.1073/pnas.1606915113 (2016). Google Scholar

55. 

Y. Li, Z. Chen and S. Ge, “Sonoporation: underlying mechanisms and applications in cellular regulation,” BIO Integr., 2 (1), 29 –36 https://doi.org/10.15212/bioi-2020-0028 (2021). Google Scholar

56. 

A. Bouakaz, A. Zeghimi, A. A. Doinikov, “Sonoporation: concept and mechanisms,” Therapeutic Ultrasound, 175 –189 Springer, Cham (2016). Google Scholar

57. 

Z. Qiu et al., “The mechanosensitive ion channel Piezo1 significantly mediates in vitro ultrasonic stimulation of neurons,” iScience, 21 448 –457 https://doi.org/10.1016/j.isci.2019.10.037 (2019). Google Scholar

58. 

S. Yoo et al., “Focused ultrasound excites cortical neurons via mechanosensitive calcium accumulation and ion channel amplification,” Nat. Commun., 13 (1), 493 https://doi.org/10.1038/s41467-022-28040-1 NCAOBW 2041-1723 (2022). Google Scholar

59. 

B. U. Hoffman et al., “Focused ultrasound excites action potentials in mammalian peripheral neurons in part through the mechanically gated ion channel PIEZO2,” Proc. Natl. Acad. Sci., 119 (21), e2115821119 https://doi.org/10.1073/pnas.2115821119 (2022). Google Scholar

60. 

M. Plaksin, S. Shoham and E. Kimmel, “Intramembrane cavitation as a predictive bio-piezoelectric mechanism for ultrasonic brain stimulation,” Phys. Rev. X, 4 (1), 011004 PRXHAE 2160-3308 (2014). Google Scholar

61. 

B. Krasovitski et al., “Intramembrane cavitation as a unifying mechanism for ultrasound-induced bioeffects,” Proc. Natl. Acad. Sci., 108 (8), 3258 –3263 https://doi.org/10.1073/pnas.1015771108 (2011). Google Scholar

62. 

G. T. Haar, “Ultrasound bioeffects and safety,” Proc. Inst. Mech. Eng. H J. Eng. Med., 224 (2), 363 –373 https://doi.org/10.1243/09544119JEIM613 (2010). Google Scholar

63. 

J. Kubanek et al., “Ultrasound elicits behavioral responses through mechanical effects on neurons and ion channels in a simple nervous system,” J. Neurosci., 38 (12), 3081 –3091 https://doi.org/10.1523/JNEUROSCI.1458-17.2018 JNRSDS 0270-6474 (2018). Google Scholar

64. 

B. M. Gaub et al., “Neurons differentiate magnitude and location of mechanical stimuli,” Proc. Natl. Acad. Sci., 117 (2), 848 –856 https://doi.org/10.1073/pnas.1909933117 (2020). Google Scholar

65. 

S. Noimark et al., “Carbon-nanotube–PDMS composite coatings on optical fibers for all-optical ultrasound imaging,” Adv. Funct. Mater., 26 (46), 8390 –8396 https://doi.org/10.1002/adfm.201601337 AFMDC6 1616-301X (2016). Google Scholar

66. 

L. Lan et al., “A fiber optoacoustic guide with augmented reality for precision breast-conserving surgery,” Light Sci. Appl., 7 2 https://doi.org/10.1038/s41377-018-0006-0 (2018). Google Scholar

67. 

P. Lavrador et al., “Stimuli-responsive nanocomposite hydrogels for biomedical applications,” Adv. Funct. Mater., 31 (8), 2005941 https://doi.org/10.1002/adfm.202005941 AFMDC6 1616-301X (2021). Google Scholar

68. 

L. Huang et al., “A compound scaffold with uniform longitudinally oriented guidance cues and a porous sheath promotes peripheral nerve regeneration in vivo,” Acta Biomater., 68 223 –236 https://doi.org/10.1016/j.actbio.2017.12.010 (2018). Google Scholar

69. 

R. Li et al., “Heparin-poloxamer thermosensitive hydrogel loaded with bFGF and NGF enhances peripheral nerve regeneration in diabetic rats,” Biomaterials, 168 24 –37 https://doi.org/10.1016/j.biomaterials.2018.03.044 BIMADU 0142-9612 (2018). Google Scholar

70. 

A. A. Al-Majed, T. M. Brushart and T. Gordon, “Electrical stimulation accelerates and increases expression of BDNF and trkB mRNA in regenerating rat femoral motoneurons,” Eur. J. Neurosci., 12 (12), 4381 –4390 https://doi.org/10.1111/j.1460-9568.2000.01341.x EJONEI 0953-816X (2000). Google Scholar

71. 

X. Chen et al., “Three-dimensional electrical conductive scaffold from biomaterial-based carbon microfiber sponge with bioinspired coating for cell proliferation and differentiation,” Carbon, 134 174 –182 https://doi.org/10.1016/j.carbon.2018.03.064 CRBNAH 0008-6223 (2018). Google Scholar

72. 

J. L. Carvalho-de-Souza et al., “Cholesterol functionalization of gold nanoparticles enhances photoactivation of neural activity,” ACS Chem. Neurosci., 10 (3), 1478 –1487 https://doi.org/10.1021/acschemneuro.8b00486 (2018). Google Scholar

73. 

J. L. Carvalho-de-Souza et al., “Optocapacitive generation of action potentials by microsecond laser pulses of nanojoule energy,” Biophys. J., 114 (2), 283 –288 https://doi.org/10.1016/j.bpj.2017.11.018 BIOJAU 0006-3495 (2018). Google Scholar

74. 

Y. Weissler, N. Farah and S. Shoham, “Simulation of morphologically structured photo-thermal neural stimulation,” J. Neural Eng., 14 (5), 055001 https://doi.org/10.1088/1741-2552/aa7805 1741-2560 (2017). Google Scholar

75. 

Y. Lyu et al., “Semiconducting polymer nanobioconjugates for targeted photothermal activation of neurons,” J. Am. Chem. Soc., 138 (29), 9049 –9052 https://doi.org/10.1021/jacs.6b05192 JACSAT 0002-7863 (2016). Google Scholar

76. 

E. Rezayat and I. Toostani, “A review on brain stimulation using low intensity focused ultrasound,” Basic Clin. Neurosci., 7 (3), 187 https://doi.org/10.15412/J.BCN.03070303 (2016). Google Scholar

77. 

N. Zheng, Multifunctional Photoacoustic Materials for Neural Engineering, Boston University( (2023). Google Scholar

78. 

T. Şen, O. Tüfekçioğlu and Y. Koza, “Mechanical index,” Anatolian J. Cardiol., 15 (4), 334 https://doi.org/10.5152/akd.2015.6061 (2015). Google Scholar

79. 

H. A. Tamaddoni et al., “Acoustic methods for increasing the cavitation initiation pressure threshold,” IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 65 (11), 2012 –2019 https://doi.org/10.1109/TUFFC.2018.2867793 ITUCER 0885-3010 (2018). Google Scholar

80. 

N. McDannold et al., “MRI investigation of the threshold for thermally induced blood–brain barrier disruption and brain tissue damage in the rabbit brain,” Magn. Reson. Med. Off. J. Int. Soc. Magn. Reson. Med., 51 (5), 913 –923 https://doi.org/10.1002/mrm.20060 (2004). Google Scholar

81. 

P. P. Ye, J. R. Brown and K. B. Pauly, “Frequency dependence of ultrasound neurostimulation in the mouse brain,” Ultrasound Med. Biol., 42 (7), 1512 –1530 https://doi.org/10.1016/j.ultrasmedbio.2016.02.012 USMBA3 0301-5629 (2016). Google Scholar

82. 

Z. Cheng et al., “High resolution ultrasonic neural modulation observed via in vivo two-photon calcium imaging,” Brain Stimul., 15 (1), 190 –196 https://doi.org/10.1016/j.brs.2021.12.005 (2022). Google Scholar

Biography

Guangxing Wang received his MS degree in optics engineering from Fujian Normal University, China, in 2020. He is working toward PhD at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. His research interests include the optogenetics and the development and applications of nonlinear optical microscopy in biological and biomedical research.

Yuying Zhou received her bachelor’s degree in food quality and safety from Fujian Medical University, China, in 2021. She is working toward MS degree at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. Her research interests include the gene therapy and the development and application of drug delivery systems.

Chunhui Yu received her bachelor’s degree from Southwest University of Science and Technology in 2022. She is currently pursuing a master’s degree at the Center for Molecular Imaging and Translational Medicine, School of Public Health, Xiamen University. Her research interests include brain science, optogenetics, ischemic stroke, neural circuits, and molecular imaging applications.

Qiong Yang received his bachelor’s degree in sanitary inspection and quarantine from Wuhan University of Science and Technology, China, in 2021. He is working toward MS degree at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. Her research interests include prevention and diagnosis of ischemic stroke.

Lin Chen graduated from Wuhan University of Science and Technology. She is working toward MS degree at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. Her research interests are the development and application of new multifunctional hydrogel materials in molecular imaging and biomedical research.

Shuting Ling is working toward her master’s degree at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. Her research primarily revolves around imaging and mechanisms associated with neurovascular diseases.

Pengyu Chen received her bachelor’s degree in food quality and safety from Fujian Medical University, China, in 2021. She is working toward master’s degree at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. Her research interests include the study of embryonic development and imaging of the embryonic brain and yolk sac.

Jiwei Xing received his bachelor’s degree in medicine from Chengdu Medical College, China, in 2022. He is working toward MS degree at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. His research interests include hydrogel synthesis, brain atrophy during aging, and artificial intelligence processing of medical images.

Huiling Wu received her bachelor’s degree in health inspection and quarantine from Fujian Medical University, China, in 2022. She is working toward MS degree at the Center for Molecular Imaging and Translational Medicine School of Public Health, Xiamen University, China. Her research interests include the optical imaging, atopic dermatitis, and the Neuropathic pain.

Qingliang Zhao is an associate professor in Xiamen University and PI of Multimodal Optical and Molecular Imaging Laboratory (MOMIL: www.zhaoqlab.com). He received his PhD from Shanghai Jiao Tong University. His research interests are building optical imaging system, molecular imaging, and optogenetics tools that can investigate the complex behavior of hemodynamic changes, biomolecules, and the interface of the microvessel and nervous systems in vivo.

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.
Guangxing Wang, Yuying Zhou, Chunhui Yu, Qiong Yang, Lin Chen, Shuting Ling, Pengyu Chen, Jiwei Xing, Huiling Wu, and Qingliang Zhao "Intravital photoacoustic brain stimulation with high-precision," Journal of Biomedical Optics 29(S1), S11520 (7 February 2024). https://doi.org/10.1117/1.JBO.29.S1.S11520
Received: 31 August 2023; Accepted: 4 January 2024; Published: 7 February 2024
Advertisement
Advertisement
KEYWORDS
Photoacoustic spectroscopy

Brain stimulation

Neurons

Ultrasonography

Brain

Acoustic waves

Light absorption

Back to Top