You are on page 1of 42

Nano Research 2018, 11(10): 5065–5106

https://doi.org/10.1007/s12274-018-2127-4

Nano functional neural interfaces


Yongchen Wang1,§, Hanlin Zhu2,§, Huiran Yang3,4,§, Aaron D. Argall5, Lan Luan2, Chong Xie2 (), and
Liang Guo3,6 ()

1
Department of Biomedical Engineering, The Ohio State University, Columbus 43210, USA
2
Department of Biomedical Engineering, The University of Texas at Austin, Austin 78712, USA
3
Department of Electrical and Computer Engineering, The Ohio State University, Columbus 43210, USA
4
Key Laboratory of Flexible Electronics and Institute of Advanced Materials, Jiangsu National Synergetic Innovation Center for Advanced
Materials, Nanjing Tech University, Nanjing 211816, China
5
Biomedical Sciences Graduate Program, The Ohio State University, Columbus 43210, USA
6
Department of Neuroscience, The Ohio State University, Columbus 43210, USA
§
Yongchen Wang, Hanlin Zhu and Huiran Yang contributed equally to this work.

Received: 15 March 2018 ABSTRACT


Revised: 5 June 2018 Engineered functional neural interfaces (fNIs) serve as essential abiotic–biotic
Accepted: 6 June 2018 transducers between an engineered system and the nervous system. They convert
external physical stimuli to cellular signals in stimulation mode or read out
© Tsinghua University Press biological processes in recording mode. Information can be exchanged using
and Springer-Verlag GmbH electricity, light, magnetic fields, mechanical forces, heat, or chemical signals.
Germany, part of Springer fNIs have found applications for studying processes in neural circuits from cell
Nature 2018 cultures to organs to whole organisms. fNI-facilitated signal transduction schemes,
coupled with easily manipulable and observable external physical signals, have
KEYWORDS attracted considerable attention in recent years. This enticing field is rapidly
neural interface, evolving toward miniaturization and biomimicry to achieve long-term interface
neurotechnology, stability with great signal transduction efficiency. Not only has a new generation
nanoelectrode, of neuroelectrodes been invented, but the use of advanced fNIs that explore other
nanomaterial, physical modalities of neuromodulation and recording has begun to increase.
neural recording, This review covers these exciting developments and applications of fNIs that
neural stimulation rely on nanoelectrodes, nanotransducers, or bionanotransducers to establish an
interface with the nervous system. These nano fNIs are promising in offering a
high spatial resolution, high target specificity, and high communication bandwidth
by allowing for a high density and count of signal channels with minimum
material volume and area to dramatically improve the chronic integration of
the fNI to the target neural tissue. Such demanding advances in nano fNIs will
greatly facilitate new opportunities not only for studying basic neuroscience but
also for diagnosing and treating various neurological diseases.

Address correspondence to Liang Guo, guo.725@osu.edu; Chong Xie, chongxie@utexas.edu


5066 Nano Res. 2018, 11(10): 5065–5106

1 Introduction facilitate new opportunities not only for studying basic


neuroscience but also for diagnosing and treating
1.1 What are functional neural interfaces? various neurological diseases.

Engineered functional neural interfaces (fNIs) serve 1.3 Why nano?


as essential abiotic-biotic transducers between an
engineered system and the nervous system. They Conventional electrode-based neurotechnologies
convert external physical stimuli to cellular signals in are facing two major hurdles: (1) susceptibility of
stimulation mode or read out biological processes in the abiotic–biotic interface to immune responses and
recording mode. Information can be exchanged using (2) communication inefficiency at the abiotic–biotic
electricity, light, magnetic fields, mechanical forces, interface [2]. Nano fNIs are compelling in offering
heat, or chemical signals. fNIs have found applications more effective solutions to both of these two aspects.
for studying processes in neural circuits from cell 1.3.1 Chronic stability
cultures to organs to whole organisms. fNI-facilitated
signal transduction schemes, coupled with easily Even though many electrode-based neurotechnologies
have made great strides during the preceding few
manipulable and observable external physical signals,
decades in proving their feasibility in treating and
have attracted considerable attention in recent years.
restoring impaired neural functions, their clinical
This enticing field is rapidly evolving toward
potential is severely restricted by issues in the
miniaturization and biomimicry to achieve long-term
integration of the neural interface within the complex
interface stability with great signal transduction
tissue environment. Not only do these mechanically
efficiency. This review covers the developments and
and chemically distinct neural interfaces cause
applications of fNIs that rely on nanoelectrodes,
significant infection, but the communication at the
nanotransducers, or bionanotransducers to establish
electrode-tissue interface is significantly diminished
an interface with the nervous system.
as the implant is isolated by fibrosis over time as a
1.2 Why are fNIs important? consequence of the foreign body reactions [3]. This
discovery of a fibrotic encapsulation developing around
In the past decade, the field of fNIs has experienced a the implanted neuroelectrodes over a short time
dramatic revolution. The once electrical-engineering window of a few months [4–7], which physically
concentrated field has evolved to a new stage that screens the electrical sensors from accessing to the
has absorbed an ever-large research population and target neurons, has largely shaped the thinking and
ever-diverse multidisciplinary approaches. Not only practice in the field in the past decade. The resulting
a new generation of neuroelectrodes has been invented, new concepts in neural interfacing have motivated
but the use of advanced fNIs that explore other physical both the development of a new generation of minia-
modalities of neuromodulation and recording has turized neuroelectrodes [2, 8–10] and the exploration
begun to increase, partially stimulated by the great of alternative approaches that feature minimum or
success of optogenetics [1]. This new stage is facilitated even no invasiveness. The reduction of the footprint
by advocations and funding support of brain-related of neural implants down to the nanoscale to make
research across the globe. Specifically, in the USA, them less “sensible” to the host tissue environment
the Brain Research through Advancing Innovative has proven to dramatically improve the chronic
Neurotechnologies (BRAIN) and Stimulating Peripheral stability of the neural interfaces [8–10]. Alternatively, to
Activity to Relieve Conditions (SPARC) initiatives mitigate the problems associated with the conventional
aim to significantly promote brain and bioelectric electrode-based approaches, such as the requirement
medicine research by accelerating the development for implantation of the bulky interface into the
and application of novel and paradigm-shifting immediate target neural tissue [11], nonspecific and
neurotechnologies, among which fNIs are a major variable activation, bio-fouling, motion artifacts, and
focus. Such demanding advances in fNIs will greatly tissue damage [3], remotely controlled approaches

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5067

that leverage nanotransducers or bionanotransducers frequently firing neurons. Unavoidable invasiveness


in the pursuit of minimum invasiveness, high spatial of implanted electrodes greatly restricts the number
resolution, and cell-type specificity have drawn and density of electrodes that can be implanted into
unprecedented attention. These alternative nano fNIs the brain, which leads to sparse sampling of the neural
hold great potentials for a better integration into the circuitry. Furthermore, conventional neuroelectrodes
target neural tissue at the tissue and cell levels. typically fail to provide consistently stable, high-quality
neural recordings over both the short and long term
1.3.2 Functional efficiency
[4, 20, 21]. In timescales as short as a few hours,
Realizing neural-realistic communication while substantial changes to the recording conditions often
minimizing side-effects to neural circuits is of great occur due to micro-movements of the implanted
interest in the field. It requires developing fNIs that electrodes with respect to the brain tissue [22–24].
comply with the biological mechanism of neural Over weeks to months, deterioration in recording
signaling. Unfortunately, signal transduction at the efficacy and fidelity is caused by biotic and abiotic
macro electrode–tissue interface is inefficient and can failures [25–28], including sustained foreign body
lead to tissue injuries due to mismatched communi- reactions at the tissue–probe interface such as neural
cation mechanisms between the electronics and the degeneration, reoccurring blood leakage in capillaries,
target tissue [12–14]. Neuronal activity is modulated and glial scar formation [5, 22, 29–32]. Recently, there
by ion channel-mediated action potential signaling, has been an increasing interest in taking advantage
and, thus, ion channels naturally become the direct of nano and micro technologies to address the
target of physical stimuli. Working at a scale close to aforementioned challenges in neuroelectrodes.
the dimensions of single ion channels, nano fNIs are
2.1.1 Nanostructure-enabled intracellular access
promising in offering a high spatial resolution, high
target specificity, and high communication bandwidth Two major types of electrophysiological recording
by allowing for a high density and count of signal methods, intracellular and extracellular, have been
channels with minimum material volume and area to developed to measure action potentials with com-
minimize tissue volumetric displacement and foreign plementary capabilities. Traditional intracellular
body reactions [15, 16]. recording methods such as the whole-cell patch clamp
technique requires rupturing a portion of the plasma
membrane to access the cell interior directly [33].
2 Nano fNIs for neural recording
Whole-cell patch clamping (recording tip diameter
2.1 Nanoelectrodes commonly ranging from close to one micron [34] to
several micrometers [35]) is the most sensitive method
Detecting the complex and dynamic activities of the to record electrophysiological events in neurons, but is
nervous system requires precise measurements of highly invasive and technically difficult to implement,
its basic functional units—neurons. Neuroelectrodes which precludes long-term or large-scale recordings.
provide one of the most useful neurotechnologies by On the other hand, extracellular recording methods
allowing for time-resolved electrical detection of neural such as multi-electrode arrays utilize micropatterned
activities and direct stimulation of neural tissues. electrodes to afford long-term and multiplexed in
Therefore, pushing the limits of electrophysiological vitro measurements [36–38]. However, extracellular
recording and stimulation is of great scientific recording suffers significantly in signal strength
and clinical interest [17–19]. Despite important and and quality. This has, therefore, resulted in the need
unique capabilities, conventional neuroelectrodes have for electrophysiological methods that combine the
significant limitations. Intrinsically, electrophysiological advantages of both intracellular and extracellular
recordings of action potentials (spikes) provide recording methods. In the past several years, there
insufficient information to establish a definite correla- have been developments that aim at achieving
tion with individual neurons and often bias towards intracellular recordings with extracellular nano-

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5068 Nano Res. 2018, 11(10): 5065–5106

electrodes or transistors while possibly allowing for a more effective and stable intracellular access to
advantages of minimal invasiveness and easy scalability. cultured cardiomyocytes. Robinson et al. fabricated
Although some structures reviewed in this section fall vertical silicon nanowire arrays that enabled intra-
into microscales instead, their working mechanisms cellular access to both record from and stimulate
rely on nanoscale structures or interactions. neurons (Figs. 1(c) and 1(d)) [43]. Abbott et al. further
It has been shown in multiple works that micro- extended this strategy by fabricating vertical nano-
and nanostructures can promote tighter contacts at electrodes on a complementary metal–oxide–
the cell–electrode interface, which enhance recording semiconductor (CMOS) multiplexer, so that an array
outcomes. Hai et al. pioneered in creating mushroom- of 32 × 32 nanoelectrodes can be simultaneously
shaped micrometer-sized gold electrode arrays that addressed [44]. A similar strategy was also reported
could detect attenuated intracellular action potentials by Liu et al. [45].
[39]. Xie et al. fabricated vertical platinum nanopillar Field effect transistor (FET) sensors, due to their
electrodes, with which they demonstrated that these sensing mechanism, do not suffer from thermal noise
electrodes formed tight junctions with cultured as the sensor size decreases, which is the major
cardiomyocytes (Figs. 1(a) and 1(b)) [40, 41]. By local limitation of passive electrodes. Tian et al. developed
electroporation, the nanopillar electrodes could FET sensors based on kinked silicon nanowires [46].
enhance the action potential recordings with more Using cultured cardiomyocytes as an in vitro model,
than ten times greater amplitudes and intracellular- they showed a clear transition from extracellular to
like waveforms. Lin et al. discovered that nanotubes intracellular recording, as the tip of the device slowly
were advantageous than nanopillars in making and penetrated the cellular membrane (Figs. 1(e)–1(j)). A
maintaining tight junctions with the cell [42]. It was free-standing version of the kinked silicon nanowire
demonstrated that iridium oxide nanotubes enabled FET sensor was created by Qing et al. [47], in which

Figure 1 Representative nanoelectrodes. (a) Scanning electron microscopy (SEM) image showing that the nanopillar electrodes were
strongly engulfed by a cell [40]. (b) Demonstration of the change in cellular membrane and corresponding recorded signal before and
after electroporation. Nanopores were formed in the plasma membrane, enabling intracellular recordings [40]. (c) SEM image of a
3-by-3 array of vertical nanowire electrodes [43]. Scale bar: 1 µm. (d) SEM image of nanowires interacting with a rat cortical neuron
[43]. (e)–(j) Nanowire FET [46]. (e)–(g) Schematic diagrams showing the recording configurations. (h)–(j) Transition of the recorded
signal from extracellular, transition period, to steady-state intracellular, as the probe slowly entered a cell. (a) and (b) adapted with
permission from Ref. [40], © Springer Nature 2012; (c) and (d) adapted with permission from Ref. [43], © Springer Nature 2012; (e)–(j)
adapted with permission from Ref. [46], © The American Association for the Advancement of Science 2010.

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5069

individual branches of the kinked nanowire could be roughness of gold from 1.7 to 22 nm, drastically
adjusted to the target-specific cell under a standard decreasing the impedance by 98% [56]. Plasma-treated
microscope. Duan et al. reported the recording of electrodes recorded signals with lower background
a full-amplitude intracellular action potential in noises and evoked local field potentials (LFPs) with
cardiomyocytes by silicon FET sensors coupled with higher amplitudes from the anterior cingulate cortex
silicon dioxide nanotubes [48]. Fu et al. pushed the in rats.
electrical detection limit further by producing a Chen et al. fabricated a carbon nanotube-based
silicon nanotube and nanowire hybrid FET electrode electrode which had a lower impedance and a six
with a recording tip size of less than 10 nm [49]. To times higher charge transfer capacity than gold
probe fine structures in the neuron, Jayant et al. took microelectrodes [57]. (More about carbon nanotube
advantage of quantum dot (QD)-coated nanopipette (CNT) electrodes are covered in section 3.1 and Figs. 2(d)
electrodes to recover full backpropagation details of and 2(e).) Kim et al. developed Au-nanotube composite
action potentials along targeted dendritic spines [50]. electrodes which reduced the impedance of gold
electrodes by 99.3% [58]. They demonstrated in vitro
2.1.2 Nanomaterials to reduce electrode impedance
extracellular recordings from mouse cortical neurons,
As the electrode size decreases to increase recording which had an average signal-to-noise ratio of 92. Kim
density or to minimize invasiveness, the impedance et al. modified electrodes with gold nanoflakes which
of the electrode increases dramatically. The thermal led to an impedance reduction from 1.15 MΩ to
noise associated with the electrode impedance becomes 26.7 kΩ at 1 kHz [59]. Kim et al. enhanced the charge
a major limitation. Nano structures and materials have storage capacity while decreasing the impedance of
great potential in significantly lowering the electrode electrodes by depositing gold nanoporous structures
impedance and extending the electrode size limitation. by dealloying Ag-Au alloy [60]. Bruggemann et al.
Owing to its highly porous structure that increases demonstrated the fabrication of vertical gold nano-
the electrochemical surface area and subsequently pillars on electrodes in an effort to achieve a higher
decreases the impedance, platinum black has long electrode surface area, which led to a decreased
been used in electrical recordings. Kim et al. improved electrode impedance [61]. Zhou et al. incorporated
the mechanical stability and electrical property of gold nanorods into a polyimide substrate to create
platinum black by depositing bioinspired adhesive a flexible thin-film microelectrode array [62]. The
polydopamine film in a layer-by-layer manner [51]. nanostructure was able to decrease the impedance by
Nanostructured porous platinum was also demon- 25 fold. Zhao et al. reported a relatively simple method
strated to increase the surface area of electrodes by to reduce the electrode impedance [63]. Electrodes
more than 30 times, thus reducing the impedance by were first electro-co-deposited Au-Pt-Cu alloy nano-
77% [52]. Weremfo et al. fabricated surface-roughened particles followed by etching away the Cu. The
platinum electrodes that had nano features and a remaining Au-Pt composite exhibited a rough surface
superior charge injection capacity comparing to with pores of different sizes. The electrode impedance
titanium nitride and similar to carbon nanotube was reduced to 4.7% of that of a bare gold electrode.
(Figs. 2(a)–2(c)) [53]. Park et al. identified an optimal Kim et al. took a hybrid approach by depositing
surface roughness factor of 233 to maximize the a layer of iridium oxide on nanoporous gold, which
performance for electrical stimulation and recording further reduced the impedance and increased the
by nanoporous platinum and yielded an impedance charge storage capacity of the electrode (Fig. 2(f)) [64].
of 0.039 Ω·cm2 and a charge injection capacity of By combining the merits of platinum gray and iridium
3 mC/cm2 (400 μs) [54]. Lee et al. deposited highly oxide, Zeng et al. [65] deposited Pt gray and IrOx on
roughened nonporous platinum film on gold tips of bare Pt in a layer-by-layer setup (Fig. 2(g)). A decent
traditional silicon electrodes and greatly decreased adhesion of IrOx was enabled by the large surface area
the interface impedance to 0.029 Ω·cm2 [55]. Chung of nanocone-shaped Pt. They achieved an impedance
et al. used CF4 plasma treatment to increase the surface value of 2.45 kΩcm2 at 1 kHz and a cathodic charge

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5070 Nano Res. 2018, 11(10): 5065–5106

Figure 2 Nanomaterials used to reduce electrode impedance. (a)–(c) Atomic force microscopy (AFM) images showing the
morphologies of smooth, 3-min roughened, and 5-min roughened platinum surfaces [53]. (d) SEM image illustrating the nanofibrous
network of multiwalled nanotube (MWNT) bundle coated on an electrode surface. Scale bar: 500 nm [66]. (e) Carbon nanotube
outperformed traditional electrode materials by having a lower impedance [66]. (f) Scanning transmission electron microscope
energy-dispersive X-ray spectroscopy (STEM-EDS) element mapping image of iridium oxide coating on a nanoporous gold electrode
[64]. (g) Morphology of IrOx/Pt gray coatings [65]. (h) Field emission scanning electron microscope (FESEM) image showing
polypyrrole (PPy) grafted at the surface of graphene oxide sheets (circled) [67]. (a)–(c) adapted with permission from Ref. [53], © the
American Chemical Society 2015; (d)–(e) adapted with permission from Ref. [66], © the American Chemical Society 2009; (f) adapted
with permission from Ref. [64], © the American Chemical Society 2016; (g) adapted with permission from Ref. [65], © Elsevier 2017;
(h) adapted with permission from Ref. [67], © Elsevier 2011.

storage capacity (CSCc) 6, 2.8, and 2.7 times higher In addition to being used as a coating material,
than those of electrodes with bare Pt, Pt gray, and graphene oxide can directly serve as the electrode
IrOx, respectively. material. Ng et al. [70] fabricated reduced graphene
Deng et al. [67] invented a one-step electrochemical oxide into disk microelectrodes of 10 μm diameter
process to deposit graphene oxide/polypyrrole com- and 60 μm pitch using nanoimprint lithography. The
posite sheet on platinum electrodes in order to increase electrodes featured a high sensitivity of 91 nA/μM to
the electrode surface roughness (Fig. 2(h)). The coated dopamine with a detection limit of 0.26 μM without
probe reduced the impedance of bare platinum using any functionalization process. The detection
electrode by 90% and increased the charge capacity capability was robust to highly resistive media,
density by more than two orders of magnitude. continuous flow, and mechanical stress.
Besides reducing the impedance, neurons seeded
2.1.3 Minimizing tissue invasiveness of neuroelectrodes
on graphene oxide doped poly(3,4-ethylene dioxy-
thiophene) (PEDOT) were found to grow longer It is commonly agreed that a long-term stable neural
neurites than those on PEDOT/polystyrene sulfonate interface that can provide reliable neural recording
(PSS). Functional biomolecules such as laminin and stimulation with minimal tissue invasiveness is
peptides could be easily bonded to the graphene of paramount importance in advancing our knowledge
surface to increase neurite outgrowth [68]. Such an of brain functions and dysfunctions, as well as
electrode demonstrated an improved sensitivity of developing next-generation neurotherapies [71]. A
151 nA/μM to dopamine compared to a glass carbon number of failure modes have been identified, and
electrode and minimized interference from ascorbic many approaches have been hypothesized and tested
acid, a competing analyte [69]. for improving the chronic stability of neural electrodes,

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5071

including but not limited to physical (dimension, and demonstrated more stable single-unit recordings
mechanical stiffness), chemical (delamination, corrosion in vivo [78]. Patel et al. fabricated 16-channel arrays
of electrodes), and biological (neuroinflammatory, of 8.4-μm diameter carbon fiber electrodes and
foreign body response) [72]. Among those highly demonstrated high-quality unit recordings for up to
interplayed failure factors, geometry and mechanical 3 months post implantation [28]. No neuronal density
mismatches between neural tissue and electrodes change and minimal to none astrocyte activation
play a critical role in determining the long-term per- were observed in post-mortem histology. Vitale et al.
formance of the electrode–tissue interface. As a result fabricated carbon nanotube fiber electrodes, which
of such mismatches, traditional neural electrodes may were softer than carbon fibers and offered smaller
exert micro-movements with respect to their host impedance and greater charge injection capacity,
tissue or apply stress. Such micro-movements not only making them more suitable for stimulation [79]. In
prevent the tracking of the same neurons as recorded addition, these softer fiber electrodes also elicited less
waveforms change accordingly, but also induce neuron chronic tissue responses.
degeneration and bleeding [73], which further cause Many efforts have also been put into making
chronic inflammatory responses, accumulation of neuroelectrodes flexible. Mercanzini et al. created a
reactive oxygen species, and eventually result in glial polyimide-based flexible probe to record LFPs, single-
scarring and accelerated electrode corrosion [74]. and multi-unit activities in the mouse cortex [80].
Seymour et al. discovered that thin electrodes (5 μm Histology results demonstrated a reduction in the
in thickness) placed as close as 25 μm from the main inflammatory response compared to rigid electrodes.
silicon electrode shank (50 μm in thickness) caused Wu et al. developed a flexible intracortical neural
noticeably fewer neuronal cell death while attracting electrode with an 8-μm thick flexible construction [81].
less immune cells [75]. Skousen et al. compared the The electrodes were shown to record neural activities
neuroinflammatory responses of two probes having an for over six weeks. Du et al. compared the chronic
identical surgical damage profile but different total tissue reactions to a novel PEDOT-polyethylene glycol
surface areas [76]. Despite the same initial damage, (PEG)-based ultrasoft microwire with conventional
the silicon lattice electrode with the smaller surface area tungsten wire electrodes [82]. At both 1 and 8 weeks
activated and attracted fewer macrophages than their post-surgery, a significant reduction in neuroin-
solid silicon counterpart. Karumbaiah et al. investigated flammatory responses was found for the soft probe
at both an acute and a chronic timescale the probe compared to its rigid counterpart. Sohal et al.
geometry-dependent inflammation by comparing the investigated long-term (26–96 weeks) foreign body
foreign body response to silicon electrode arrays responses in the rabbit cortex induced by flexible
that had disparate thicknesses [77]. A 15 μm version parylene-C-based electrodes with microwire electrodes
outperformed a 50 μm version in terms of histological as the control [83]. Less gliosis and greater neuronal
analysis and activated immune cell density around density were observed for the flexible probe. More
the probe at both 3- and 12-week checkpoints. These importantly, those effects were more prominent
studies all underscore the importance of electrode towards the long-term timescale.
geometrical dimensions for mitigating tissue reactions. It is clear that the switch from conventional rigid
Carbon fibers have recently been a popular materials, such as silicon and metals, to softer polymers,
choice to fabricate thin neuroelectrodes. Kozai et al. such as polyimide and parylene, helps mitigating tissue
demonstrated ultra-small 7-μm diameter carbon fiber reactions and promotes recording stability. However,
electrodes for high-quality extracellular recording [8]. these polymers are still orders of magnitude stiffer than
These electrodes were shown to have significantly the brain tissue [84], and therefore the mechanical
reduced foreign body responses and to record mismatch is still prominent. On the other hand, it is
single-unit activities for over 4 weeks in vivo. impractical, at present, to fabricate functional devices
Guitchounts et al. employed bundles of 5-μm diameter with materials that are as soft as tissues. This
carbon nanofibers to fabricate penetrating electrodes dilemma led researchers to try different approaches

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5072 Nano Res. 2018, 11(10): 5065–5106

besides using alternative materials. Because most is to simultaneously record from a large number
neuroelectrodes are constructed in high-aspect-ratio of neurons. The nature of extracellular recordings
probe geometries, bending is the major deformation determines that the electrode must be within the
happening at the interface. Therefore, the primary close vicinity (~ 100 μm) of a neuron to precisely
goal of mitigating the electrode-tissue mismatch is to detect its activity [90]. To record from many neurons
minimize the bending stiffness Ks of the implanted inevitably requires a large array of electrodes placed
probe, which is given by Ks = Eswh3/12 for a probe on the surface of the brain or inserted into non-
with a rectangular cross-section [10], where Es is the superficial structures. Besides, recording throughout
elastic modulus of the material, w is the probe width, a region requires that all neurons are within the
and h is the probe thickness. It is clear that the probe range of a recording site and that recording sites are
geometry (thickness and width) plays an important densely packed to allow accurate spike sorting. In
role in modulating the bending stiffness. Therefore, particular, the detection range of each recording site is
nanoelectronic devices could have a unique and typically 50–100 μm [91, 92], and spike sorting requires
important impact in promoting tissue integration of an electrode spacing of 20–50 μm center-to-center
neuroelectrodes. [93, 94] to resolve more single neurons.
A series of recent work has demonstrated neural In the efforts to push limits of recording density,
probes made by nanoelectronic devices based on channel count, and single-unit yield, significant
ultrathin polymer structures (Fig. 3) [9, 10, 85–89]. The challenges arise, such as addressing individual channels
key feature of 1-μm thickness yielded ultra-flexibility and tissue displacement/injury.
that allowed for a great reduction in mechanical Guo and DeWeerth developed an effective lift-off
mismatches at the tissue–electrode interface. This method to pattern gold wires at high density on
enabled chronically stable recordings from the same stretchable substrates [95]. Based on this technique,
neurons over multiple months. high-density compliant and stretchable electrode arrays
were fabricated and demonstrated by interfacing with
2.1.4 High-density neural recording
muscular tissue [84]. Khodagholy et al. demonstrated
One of the primary challenges of neurotechnologies the use of organic transistors to record neural signals

Figure 3 Increased flexibility improves the tissue-electrode interface. (a) Photoacoustic images of two flexible CNT electrodes inserted
in the brain [89]. (b) Micrograph of a three-dimensional (3D) mesh nanoelectronic brain probe suspended in buffer with a cylindrical
shape [9]. (c) Chronic tracking of two detected neurons over a course of 4 months post-surgery [10]. Scale bar: vertical, 200 µV;
horizontal, 1 ms. (d) Nanoelectronic probe as fabricated on a substrate [10]. Scale bar: 50 µm. (e) Two-photon imaging demonstrating
little dislocation between neurons and an ultra-flexible electrode for one month [10]. Scale bar: 50 µm. (f) Confocal micrograph of an
immunochemically-labeled brain slice 5 months post-surgery, showing neurons surrounding the probe with minimal microglia activation
[10]. Neurons labeled in orange, electrode labeled as a green rectangle, and microglia indicated by white arrows. Scale bar: 50 µm.
(a) adapted with permission from Ref. [89], © the American Chemical Society 2013; (b) adapted with permission from Ref. [9], © Springer
Nature 2015; (c)–(f) adapted with permission from Ref. [10], © the American Association for the Advancement of Science 2017.

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5073

in vivo [96]. The transistors were made of PEDOT on active recording devices with buffer amplifiers placed
a highly flexible parylene substrate with a total closely to individual electrodes were also developed
thickness of 4 μm. These devices are capable of for high-density penetrating electrodes. Lopez et al.
amplifying and multiplexing signals locally and patterned 455 electrodes on a shank width of 100 μm,
potentially support high-density surface recordings. among which 52 could simultaneously record, using
Viventi et al. fabricated a high-density electrode array 180 nm CMOS technology [102]. Lopez et al. further
multiplexed by silicon nanomembrane transistors on developed this technology and achieved 384 con-
a flexible polyimide substrate [97]. LFP recording figurable channels out of 966 electrodes on a 70 μm
on the brain surface was demonstrated in an animal wide shank with 130 nm CMOS technology [103].
seizure model. Recordings with two of such active CMOS probes
There are also significant recent efforts on making were demonstrated in freely moving rats (Fig. 4(f))
penetrating neuroelectrodes with nanofabrication. [104]. A total of 700 single units were recorded from
Du et al. nanofabricated a 64-channel probe with five brain structures, demonstrating the capability of
interconnect width and spacing of 290 nm and electrode acquiring large-scale activities of hundreds of neurons.
pitch from 28 to 40 μm [91]. They found that only High-density nanosensors also enabled mapping
50% of all identified putative thalamic neurons were of electrochemical activities of neurons at superior
detectable with electrodes separated more than 40 μm, spatial resolution. With more than 20,000 sensors per
which justifies the importance of the electrode array cell, Kruss et al. [105] fabricated single-wall nanotube-
density. Similarly, Rios et al. nanofabricated electrodes based sensors that detected dopamine release at 100 ms
with features as small as 300 nm to enable high- resolution. They were able to investigate how the
density 3D recordings [98]. They achieved a minimum chemical communication of neurons was affected by
inter-electrode distance of 16 μm, totaling 1,024 cell morphology and developed models to optimize
channels (64 channels per shank) and covering a tissue their probe design based on the spatially preserved
volume of 0.6 mm3. Single units were found to show data [106].
up on more than six channels, allowing for neuron For a detailed review of high-density neural
trilateration. Unique layer-specific field potential
electrode array architectures and their respective merits
signatures across the whole hippocampus facilitated
and challenges, readers are referred to Refs. [107, 108].
the determination of layer boundaries, which was
subsequently validated by histology results. An even 2.2 Nanotransducer-assisted functional neural
denser array was created by Scholvin et al. with imaging
200 nm wiring width and 400 nm spacing [99]. 1,000
electrodes (200 channels per shank) were patterned Imaging the electrical activity of neurons and neural
on silicon probes with 11 μm electrode separation networks is of fundamental importance in under-
(Figs. 4(a) and 4(b)). Typically, a recorded unit could standing their physiological functions and treating
manifest on more than 10 channels. Obien et al. their pathological dysfunctions. Their electrical activity
fabricated a 128-channel electrode array with on-chip can be recorded through the electrodes in a patch clamp
micro light emitting diode (μLED) for high-resolution or microelectrode array. These electrical recording
electrophysiology coupled with optogenetics [100]. techniques are invasive and incapable of recording
Wei et al. nanofabricated high- density electrode arrays neural activity from a large region of interest, which
on an ultra-flexible structure (Figs. 4(c)–4(e)) [101]. has promoted the optical imaging approaches. As
The interconnect traces had a width of 200 nm and calcium ions (Ca2+) are a critical indicator of neuronal
a pitch of 400 nm, addressing electrodes spaced as activity, Ca2+ imaging is a powerful tool to study
small as 20 μm. The overall cross-section area of the neural activity [109]. However, the Ca2+ dynamics is
probe was as small as 10 μm2, which minimized chronic slower than that of the actual action potential, and
tissue reactions. only suprathreshold neural activities can be recorded
In addition to the passive approach, CMOS-based [110]. Thus, directly imaging the transmembrane

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5074 Nano Res. 2018, 11(10): 5065–5106

Figure 4 Nanofabricated high-density neural electrodes. (a) Individual high-density probe with 200 electrode channels [99]. The
electrodes were spaced by 11 µm in a one-dimensional (1D) arrangement. (b) Recorded spikes using the probe in (a). (c)–(e)
High-density ultra-flexible probes with 20 µm electrode pitch and 800 nm thickness [101]. (f) Schematic diagram and microscope image
of an active CMOS high-density probe [104]. (a) and (b) adapted with permission from Ref. [99], © IEEE 2016; (c)–(e) adapted with
permission from Ref. [101], © Wiley-VCH 2018; (f) adapted with permission from Ref. [104], © Springer Nature 2017.

potential of neurons with high spatiotemporal (micro- potential of an action potential could result in such
meter and sub-millisecond) resolutions, sensitivity photoluminescence quenching and red-shift in type-II
up to subthreshold activity, and versatility for both QDs and that QDs would have a higher sensitivity
excitation and inhibition is greatly desired. Although compared to VSDs [114]. In a follow-up work, the use
the voltage-sensitive dyes (VSDs) for voltage imaging of type-I and quasi-type-II QDs to image a voltage
(discussed in detail below) have a high spatiotemporal resembling an action potential with millisecond
resolution and capability of imaging a large region temporal resolution was experimentally confirmed
of interest, they are constrained by sensitivity, [111]. In this work, it was also claimed that the
photobleaching, and phototoxicity [111]. Thus, semi- photoluminescence quenching by QDs was attributed
conducting nanotransducers, especially QDs, with to an increase of ionized QDs and that quasi-type-II
higher photoluminescence intensity and photostability QDs had a higher sensitivity than type-I QDs. In a
have attracted attention for potential application to more recent theoretical study, modeling showed that
voltage sensing in neurons. type-I and type-II semiconducting nanorods could
Due to quantum confinement, QDs can be excited have an even higher sensitivity than QDs, while the
by light, creating excitons, pairs of separated negatively sensitivity of type-II nanorods was higher than that
charged electrons and positively charged holes [112]. of type-I nanorods [115]. However, the nanorods
An electric field applied to the excited QDs leads to a needed to penetrate the plasma membrane vertically
quantum-confined Stark effect, decreasing the energy and symmetrically, which was technically challenging.
of electrons and holes [113]. As a result, their Up to now, the QD- or semiconducting nanorod-
photoluminescence intensity is quenched, and the assisted voltage imaging has not been biologically
emission peak is red-shifted and broadened. This validated. The physiological scenario is far more
property well aligns with the needs for optical voltage complicated than the theoretical simulation. Once
imaging. A theoretical study showed that the membrane tested in neurons, there are still major practical

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5075

challenges associated with their voltage sensitivity, information processing in the nervous system [117,
placement, and biocompatibility. Voltage imaging 118]. After decades of development and optimization,
requires the nanosensors to be placed in the vicinity VSDI exhibits facile delivery into living preparations
of or within the plasma membrane. The QDs delivered for long-term observation, from dissociated neurons
to the plasma membrane are, however, rapidly to frogs, rats, and nonhuman primates [119–121].
internalized, and the placement of QDs is dynamically Typical VSDs possess an amphiphilic structure, in
changing. Considering the fast voltage dynamics which the hydrophobic components work as anchors
compared to the endocytosis dynamics, this should into the cellular membrane, and the hydrophilic
not affect a single measurement within a short period components, mostly chromophores, which array per-
of time but will require repetitive dosing for separate pendicularly to the cellular membrane. The intensity
measurements, which will worsen the cytotoxicity changes of the fluorescent signal correlate to the
of QDs. Thus, exploring nanotransducers that can transmembrane voltage changes, and the neural activity
convert voltage to an observable and readable signal can be measured accordingly. The most common
and meanwhile can serve as stable and biocompatible imaging mechanism is redistribution (Fig. 5(a)). A
interfaces is essential for the development of change in the transmembrane potential causes the
nanotransducer-assisted voltage imaging in neurons. charged chromophores to move in or out of the cell,
leading to a regional concentration change of the
2.3 Bionanotransducer-enabled functional neural chromophores with a corresponding change in the
imaging fluorescence intensity. Another mechanism is called
reorientation (Fig. 5(a)), which is determined by
Compared to electrode-based neural recording
the interaction of the intermolecular electric fields.
approaches, optical techniques offer complementary, Additionally, electrical modulation of the electronic
yet compelling, advantages for imaging the tran- structure and fluorescent resonance energy transfer
smembrane potentials, including less invasiveness (FRET) can cause fluorescence changes in response to
and superior spatial resolution [116]. Optical recording changes of the transmembrane potential (Fig. 5(b)).
or imaging can capture both the electrical activities of In addition to the commonly used VSDs
each neuron in the circuit and map with micrometer (aminonaphthylethenylpyridinium (ANEP) and Rina
resolution and sub-millisecond precision. Hildesheim (RH) families) [122], QD-based approaches,
Nanobiotechnologies for optical neural imaging as mentioned above, provide an alternative opportunity.
are mostly based on genetically encoded proteins, for Functional modifications using biomolecules can
their remarkable biocompatibility and optical properties, both reduce the cytotoxicity of QDs and provide
fast response dynamics, high effective sensitivity, and functional targeting sites to the neuronal membrane
easy functional modification. Protein-based optical for voltage imaging [114]. Recently, Nag et al. reported
sensors can be used to measure transmembrane a QD-peptide-fullerene nanobioconjugate for imaging
potentials in single cells, tissues, and living animals membrane potentials in living cells [123]. This
through detection of the exhibited fluorescent signals. alanine/leucine-rich peptide was designed to form a
A number of optical imaging methods based on a helix to promote its membrane insertion. It also could
variety of mechanisms have been explored for the append the fullerene component at discrete fixed
benefits of neuroscience research. distances for the signal detection and facilitate energy
transfer via tunneling. The imaging of PC-12 cells
2.3.1 Voltage-sensitive dye imaging
showed a 20- to 40-fold improvement in F/F with
Traditional voltage-sensitive dye imaging (VSDI) is no sacrifice in responsivity. Thus, protein-modified
based on a fluorescent change of small molecular dyes nanotransducers also play important roles in specific
when subjected to a change in a local electric field. targeting to the cellular membrane, enhancement of
For its good spatial (up to 20 m) and temporal (up biocompatibility, and connection between two FRET
to tens of microseconds) resolutions, this approach components, with minimum effect on the fluorescent
offers a great potential in the visualization of the signal.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5076 Nano Res. 2018, 11(10): 5065–5106

Figure 5 Bionanotransducer-enabled functional neural imaging. (a) and (b) Mechanisms of VSD imaging: redistribution ((a), left),
reorientation ((a), right), and FRET (b). (c) GECIs based on single fluorophore and FRET mechanisms. After binding of Ca2+ to GCaMP,
conformational changes induce an increase in emission (left) or a change in the ratio of emission intensities (YFP/CFP) [124, 125].
(d) Mechanisms of GEVIs. Left, single FP-based VSFPs exhibit fluorescent quenching upon membrane depolarization; right,
FRET-based VSFP-Butterfly family has a fused four-transmembrane-segment voltage-sensitive domain between two FPs [126].
(e) VSFP2.3 reported transmembrane voltage transients through two different FPs [127]. (f) Extracellular electrical stimulation induced
an increase in intracellular Ca2+ concentration in myotubes with the expression of GCaMP [124]. (c) adapted with permission from
Refs. [124, 125], © Springer Nature 2001 and National Academy of Sciences 2004, respectively; (d) adapted with permission from Ref.
[126], © Springer Nature 2015; (e) adapted with permission from Ref. [127], © Springer Nature 2010; (f) adapted with permission from
Ref. [124], © Springer Nature 2001.

2.3.2 Genetically encoded fluorescent imaging other domains that in response to cellular events
undergo conformational changes such as intracellular
Neural activities are encoded by dynamic fluctuations Ca2+ concentrations, transmembrane potentials, small
of the transmembrane potential, including both metabolites, and other ions. Two major classes of
subthreshold and action potentials. To visualize the genetically encoded optical indicators have been
dynamic changes of the transmembrane voltage, which developed: One contains genetically encoded calcium
range from ~ 1 mV (e.g., a postsynaptic potential indicators (GECIs), which have a Ca2+-binding domain
caused by a single vesicle release) to over 100 mV (i.e., and a Ca2+ concentration response conformation;
an action potential) and span from 100 ms (axonal another comprises genetically encoded voltage in-
conduction) to 10 s (plateau potential), genetically dicators (GEVIs), which have a domain responsive to
encoded optical indicators offer a great promise. transmembrane potential changes.
Nanobiotechnological engineering approaches have Genetically encoded calcium indicators: Developed
been used to fuse the fluorescent chromophores with since 1997 [128], GECIs monitor changes to the

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5077

intracellular Ca2+ concentration caused by action the excitation and emission. A number of improved
potentials. As a ubiquitous second messenger, Ca2+ is variants of the original blue, cyan, yellow, and
of great significance in all aspects of cellular physiology. red FPs [126] have been developed to improve the
Triggered by an action potential, an increase in its brightness, photostability, tissue penetration, and
intracellular concentration, from 20–100 nM (initial signal-to-noise ratio.
concentration) to 5–10 M (peak concentration), causes To avoid interferential interactions of CaM with
a change to the fluorescence [129]. Briefly, the design endogenous binding factors and improve the detection
of GECIs involves the fusion of two parts: a naturally precision, two effective approaches are pursued: One
evolved Ca2+-binding protein component that undergoes is the replacement of CaM with troponin C variants,
a conformational state transition in response to which is the Ca2+ binding protein in muscle cells and
Ca2+ binding (calmodulin (CaM) [128] or troponin-C does not have an endogenous binding site in neurons
(TnC) [130]) and a reporter component based on a [130, 135]; the other is the modification of binding
conformation-sensitive fluorescent protein (FP). interference. Palmer et al. reengineered the binding
Two types of signals are recorded according to interface between CaM and a target peptide to generate
the type of reporter component (Fig. 5(c)). A simple selective and specific binding pairs that could be no
fluorescence signal is collected through the confor- longer perturbed by large excesses of native CaM, and
mational transition of a single FP component [124]. A the Ca2+ detection sensitivity turned over a 100-fold
pair of FPs exhibits FRET, which could provide a range at 0.6–160 M [136].
measure of the Ca2+ concentration by the ratiometric In addition, the optimization of GECIs focuses on
fluorescence signal of the two FPs [128]. The FRET their modulation for specific applications and better
signal is a better alternative due to its independence optical properties, such as optimization on the Ca2+
from probe concentration, low excitation light intensity, binding constant [131, 137], subcellular targeting [138],
and low absorption in the optical path. For example, brightness, and red-shifting of the fluorescent indicators
as a typical GECI, yellow Cameleon (YC) 3.60 has a for large penetration and good resolution [139].
pair of enhanced cyan FP (ECFP) as the donor and GECIs can be used in combination with two-photon
Venus protein FP with circularly permuted confor- imaging to achieve improved measurements in a highly
mation (cpYFP) as the acceptor. In the presence of scattering medium with an increased signal-to-noise
Ca2+, its intramolecular conformation changes lead to ratio without the need for averaging. Mank et al.
a reduced spatial distance and illuminate the Venus generated a GECI as TN-XXL for two-photon ratio-
protein, and the ratio change of YFP/CFP is up to six metric imaging in the visual cortex. They rearranged
folds with a high Ca2+ dynamics up to 0.25 M [125]. the Ca2+ sensing moiety TnC within the indicator and
The mostly optimized GECIs are single-wavelength mutagenesis of selected amino acid to increase
green indicators based on the original GCaMP sensor. overall signal strength and sensitivity in the low Ca2+
The reporter domain of GCaMPs is a circularly regime [140].
permuted enhanced green FP (EGFP), which is flanked GECI fluorescence imaging is a good proxy to
by the Ca2+ binding protein CaM and the CaM-binding record average action potential changes, but it has an
peptide M13. In the presence of Ca2+, CaM-M13 inherent inadequacy. Because Ca2+ transients are
interactions lead to conformational changes and an 100–1,000 fold slower than the underlying electrical
increase in fluorescent emission [124, 131]. A number waveforms, most subthreshold changes of tran-
of engineered variants of GCaMP have been reported, smembrane potentials cannot elicit a change in the
and these indicators are composed of constituent Ca2+ concentration, and the Ca2+ dynamics is confounded
molecules to achieve specific manipulation of sensor by the complicated interactions between different
applications [132–134]. Ca2+ sources and intrinsic or extrinsic Ca2+ buffers.
Since the first GECI was reported with the chromo- Hence, weak and brief changes of the transmembrane
phore green FP (GFP), the modulation of indicator potentials may not be recorded by the Ca2+ indicators
color has been well explored. Directed mutation of [110, 141].
the side-chains comprising the chromophore can tune GEVIs: GEVIs may be preferred over GECIs, for

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5078 Nano Res. 2018, 11(10): 5065–5106

their capability of directly reporting both synaptic a voltage sensing domain of a K+ channel and a pair
and action potentials as well as capturing the entire of cyan and yellow emission mutants of FPs. Although
voltage dynamics. Several different approaches to this generation of VSPFs could optically report changes
build GEVIs have been developed since 1997. Most in the transmembrane potential, their application to
GEVIs consist of two components: a voltage-sensitive functional imaging of mammalian neurons was limited
domain from an ion channel as the voltage sensor by their low targeting capability to the membrane
and a component that binds in the plasma membrane [150]. The second generation of VSPF (VSPF2) was
and experiences the voltage changes. based on Ciona intestinalis voltage sensor-containing
Same as for GECIs, two types of fluorescent signals phosphatase (Ci-VSP), exhibited better membrane
based on different response mechanisms can be targeting, and formed a big Ci-VSP GEVI family
collected: One is the single fluorescent signal from a [116]. Akemann et al. developed FRET-based VSFP2s
single fluorescent moiety [142–144], which undergoes by changing the enzyme domain of Ci-VSP with two
a significant conformational change that alters its tandem fluorescent domains, which exhibited more
spectrum; and the other is a ratiometric signal based efficient targeting to the cellular membrane and higher
on FRET between a pair of fluorescent moieties [127, responsiveness to the transmembrane potentials [127].
145], which are both involved in the molecular The variant VSFP2.3 was developed as a FRET-based
interactions and motions, leading to ratio changes indicator possessing the capacity of imaging both
in fluorescence intensities (Fig. 5(d)). Signals from a spontaneous action and synaptic potentials in neurons
single FP have limited sensitivity, while the ratiometric [151]. The indicator VSPF3.1 was the fusion of Ci-VSP
FRET signals can significantly reduce the motion and and a single red-shift protein, which offered the
blood flow effects in complex environments. advantages of a red-shifted spectrum and relatively
Several GEVIs based on different fluorescent fast overall kinetics [152]. The VSPF-Butterfly series,
chromophore families have been developed. For with a sandwiched structure between two FPs and
example, the fluorescent Shaker (FlaSh) and the sodium a middle voltage-sensitive domain, showed a
channel protein-based activity reporting construct subthreshold detection range and fast kinetics for
(SPARC) are based on simple fluorescent signal single-cell synaptic responses in applications from
detection, while the voltage-sensitive FP (VSFP) uses single neurons to the brain [153].
an FP pair based on FRET [146–148]. Besides of FP-based GEVIs, microbial rhodopsin
Siegel et al. reported an early attempt on FlaSh could also be used as the fluorescent voltage reporter
[146]. They constructed a GFP-Shaker fusion protein, [154, 155]. This retinal chromophore shows very weak
using a nonconducting mutant of a voltage-gated K+ near-infrared fluorescence, while upon light-driven
channel as the voltage-responsive site and an FP transport of protons, a change in the chromophore
inserted into the K+ channel protein as a reporter. This emission was induced. Comparing to FP-based
GEVI had a maximal fractional fluorescent change of GEVIs, GEVIs based on the microbial rhodopsin have
5.1%. Guerrero et al. expanded the FlaSh GEVIs by a sub-millisecond response time and better voltage
modifying kinetics, dynamic range, and color. The sensitivity [154–156]. GEVIs based on non-pumping
improved folding enhanced the detection sensitivity, mutants of Archaerhodopsin 3 (Arch) have voltage
and the availability of different FP colors promoted sensitivities between 30%–90% per 100 mV and
wide adoptions [149]. The SPARC was developed half-maximal response times between 50 s and 1.1 ms
with a strategy to insert a GFP molecule into a rat at room temperature [156]. However, Arch-derived
muscle Na+ ion channel subunit, which exhibited rapid GEVIs exhibit weak fluorescence, more than 30-fold
response kinetics without significant inactivation [147]. weaker than that of FPs. Meanwhile, to illuminate
As the key component in the FP-based GEVIs, VSFP microbial rhodopsin-based GEVIs, the excitation
underwent a lot of research. The first generation of laser power needs to be much higher, typically 300–
VSFPs was derived from the voltage-sensing domain 1,000 W/cm2, much more than that of FP-based GEVIs
of a K+ channel subunit [148]. This VSFP consisted of (10 W/cm2) [156]. The low brightness of rhodopsin-

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5079

based GEVIs presents a challenge for its widespread have been shown to improve the charge storage
use. capacity and charge injection limit of microelectrodes,
Current efforts on the development of GEVIs which results in more effective stimulation including
primarily focus on the optimization of existing sensor smaller voltage, greater power efficiency, and less
constructs, including the voltage range, color, brightness, tissue damage.
response kinetics, and cellular targeting properties. Carbon nanotube-based electrodes were found to
As key indicators to evaluate sensing performance, offer lower impedance and improved charge injection
response sensitivity (F/F), half-maximal response compared to platinum, making it a suitable candidate
times, signal-to-noise ratio, and excitation power all for electrical recording and stimulation. Wang et al.
need to be considered. The first generation of GEVIs first reported the use of vertically aligned multiwall
exhibited modest voltage sensitivity (< 5% F/F per carbon nanotubes to stimulate excitable cells (rat
100 mV) and slow kinetics (10–200 ms). The second hippocampus neurons) [157]. Tsang et al. reported
generation of Ci-VSP-based GEVIs had a novel GEVI the use of Au-carbon nanotube composite electrodes
(Arclight) consisting of a Ci-VSP and a super ecliptic fabricated on flexible polyimide substrate [158]. As a
pHluorin that carried the point mutation A227D [142]. result of the lower stimulation voltage and less power
Arclight A242, derived from Arclight, exhibited a consumption, the stimulation could be done wirelessly.
fluorescent intensity increase by 35% F/F per 100 mV, The resulting device forms an insect–machine interface
but the response time (10 ms) was much slower, in which the flight path of a moth could be biased by
hindering action potential detection. In another GEVI selectively stimulating one side of the moth’s body. Yi
design, called ASAP1, a circularly permuted FP was et al. developed vertically aligned carbon nanotube
inserted in an extracellular loop of a voltage-sensing electrodes on flexible substrates and used them to
domain [144]. ASAP1 exhibited high sensitivity stimulate rat sciatic nerves and record from rat spinal
(18%–29% F/F per 100 mV) and high kinetic speed nerves with a signal-to-noise ratio as high as 12.5 [159].
(2 ms) and could be used for action potential detection Jan et al. quantified the impedance and charge storage
at waveforms up to 200 Hz. capacity of iridium oxide, PEDOT, and layer-by-layer
GECIs and GEVIs have exhibited the capability of synthesized multiwall carbon nanotube [66]. Carbon
recording neuron potentials in multiple mammals nanotubes outperformed traditional electrode materials
and neuron types, but the use of them in humans by having a lower impedance and higher cathodic
seems unlikely at present because its delivery involves charge storage capacity. In addition, no sign of failure
viral vectors. The development and optimization of was observed after 300 cycles of cyclic voltammetry
genetically encoded indicators still continue. The ideal scans.
GECIs and GEVIs should process properties such as
large voltage/Ca2+ induced fluorescent changes over 3.2 Nanotransducer-assisted neural modulation
a physiological range, quick response kinetic time The emerging concept of going wireless has been
(sub-millisecond), targeting capability to different revolutionizing the contemporary technologies and
neuron types, good photostability, large brightness and reshaping the modern lifestyle. It is also driving the
signal-to-noise ratio, and red/near-infrared emission advancement of medical devices [160]. Conventional
for in vivo applications. neural modulation techniques deliver electrical signals
to the target neural population. Due to fast dissipation
3 Nano fNIs for neural stimulation and attenuation of the electrical signals through tissues,
these traditional techniques often demand invasive
3.1 Nanoelectrodes placement of the electrodes in the immediate vicinity
of the target neural regions and implantation of a
3.1.1 Microstimulation with nanomaterial coatings
pulse generator wired to the electrodes to deliver the
Nanostructured surfaces and nanomaterial coatings electrical signals. Surgical procedures cause tissue

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5080 Nano Res. 2018, 11(10): 5065–5106

damage and surgical complications [161], and chronic rationales, as the secondary signal is the ultimate
inflammation around the electrodes causes scar modulatory signal to the target neuron. For example,
tissue formation and early device failure [14]. manipulation of ion channels plays an important role
Alternative strategies to neural modulation, which in neural modulation, and there are various types of
rely on direct wireless delivery of magnetic field, ion channels, including voltage-gated, temperature-
ultrasound, and infrared light to the target neurons, gated, mechanosensitive, and light-gated. Thus, the
unnecessitate the electrodes but are limited in spatial local secondary signal needs to be generated by the
resolution and/or power efficiency. These deficiencies nanotransducers to gate the corresponding ion channels
have motivated the development of a new generation and eventually modulate the neuronal activities.
of nanotransducer-assisted wireless neural modulation Here, these diverse types of nanotransducers can be
techniques featuring minimal or even no invasiveness classified into electrotransducers, thermotransducers,
with greatly improved spatiotemporal precision and optotransducers, and mechanotransducers based on
cellular targeting specificity. the type of secondary signal they generate.
Like wireless charging, in which the electromagnetic To develop such nanotransducer-assisted wireless
energy is transmitted through the intermedium and neural modulation techniques, there are several
converted to electrical currents by an antenna in the key factors to consider, including the primary and
target battery, a primary wireless signal (e.g., optical, secondary signals, the nanotransducer’s biosafety,
magnetic, or acoustic) can be transmitted through biostability, placement, and modification, and cell
tissues and converted to a secondary local signal (e.g., modification.
electric, thermal, optical, or mechanical) by nano Primary signal: The primary signal needs to be safe
antenna-transducers to modulate the target neuron. and wirelessly transmittable through tissues to reach
These nanotransducers are essential for delivering the the target region. Although light, magnetic field, and
highly localized secondary signal with spatial precision, ultrasound can all meet these two criteria to possibly
target specificity, and improved efficiency. Due to serve as a primary signal, their interactions with tissues
their small size and novel physicochemical properties, affect their penetration depth and spatial resolution
these nanomaterial-based nanotransducers perfectly (i.e., how well the primary signal can be focused on
align with these demands. Firstly, they can be delivered the target region). Light has superior spatial resolution
to the target region via minimally-invasive local or of 10–7 m, while it can barely pass the cranium (10–6 m
intravenous injection. Secondly, their diverse energy travel depth) and only penetrate dermal tissues as
transduction schemes allow the conversion of a deep as 4 mm [163, 164]. Thus, for deep neural tissue
medically safe, tissue-penetrating primary signal to modulation, implantation of a light source is usually
a localized cell modulatory secondary signal at the required. Transcranial focused ultrasound traveled
nanotransducer–neuron interface [15]. Additionally, through the skull and modulated the cortical activity
their ease of surface modification and bio-conjugation 30 mm deep in human brains with a spatial resolution
can facilitate specific targeting to the targeted neuron of 10–3–10–2 m [165]. The penetration depth and spatial
population at cellular and even subcellular levels, resolution of transcranial magnetic stimulation are
dramatically improving the target specificity, selectivity, highly dependent on the coil design. For example,
and spatiotemporal resolution. the 50 coil design has a penetration depth and spatial
Over the past few decades, particularly the last resolution of 10–2 m [166]. For these primary signals,
decade, a great deal of efforts has been made to there is a tradeoff between penetration depth and
develop such nanotransducer-assisted wireless neural spatial resolution [165, 166]. Additionally, the interac-
modulation techniques. In the literature, these tions of primary signals with tissues may also generate
techniques are commonly categorized by the primary noxious heat, limiting the maximum intensity of the
wireless signals. In contrast, in the present review, primary signal.
these techniques are categorized by the secondary Secondary signal: The secondary signal needs to be
local signals to elucidate the concepts and design able to modulate neurons by gating the respective ion

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5081

channels, activating cell signaling pathways, changing nervous system. On the other hand, binding of
the plasma membrane capacitance, etc. Possible nanotransducers to the plasma membrane, receptors,
secondary signals include electric fields, heat, light, or ion channels improves the specificity and washout
and mechanical forces. Generation of the secondary resistance, but they can be rapidly internalized. The
signal should be controlled to exceed the threshold for internalized nanoparticles can cause toxicity and
cellular modulation but not to exceed the safety limit modulation inconsistency [167, 168]. Additionally, a
that can cause cell damages. Theoretical simulation complete retrieval or clearance of the nanotransducers
can be used to help assess the feasibility of a signal after the modulation can be challenging for dispersed
transduction scheme and guide the experimental nanoparticles.
design and optimization. Additionally, the temporal Nanotransducer modification: Nanotransducers are
resolution of the modulation significantly depends usually surface-modified so that their dispersity
on how fast the neurons respond to the secondary and biocompatibility can be improved. They can be
signal. immobilized on a substrate, specifically targeted to a
Nanotransducer biosafety and biostability: The nano- neural population or even to the receptors and ion
transducers need to serve as stable neural interfaces channels. However, surface modification can attenuate
within the modulation period, and they must be safe the secondary local signals and increase the distance
in the physiological environment. They should cause between the nanotransducers and the modulation
no toxicity and minimal inflammation. Moreover, target. Moreover, the surface chemistry needs to be
they should remain intact to maintain the functional well controlled, because certain surface chemistry,
interfaces and not release or leak toxic substances. even without inducing clear cytotoxicity, may cause
Their biostability can be challenging. First, certain false modulatory effects due to plasma membrane
nanotransducers are prone to degradation due to the perturbation [169].
physiological acidic pH and enzymatic activities. Cell modification: Genetically engineering certain
Second, the exogenous nanotransducers are also prone types of neurons to express the target ion channels
to clearance, so their placement is dynamic instead of enables and optimizes the cells’ response to the
static. corresponding secondary local signal. Neurons can
Nanotransducer placement and distribution: Overall, also be genetically modified to express binding sites
nanotransducers can either be in a dispersed/ for nanotransducer docking and create hallmarks
injectable form or immobilized onto a substrate or for evaluating the modulatory effect. These cell
microelectrode (Fig. 6(a)). The dispersed nanotran- modification strategies make the modulation technique
sducers can be placed extracellularly, targeted to the versatile and adaptable for different types of cells.
plasma membrane (e.g., onto the membrane, receptors, They can also improve the modulation specificity
anchoring proteins, or ion channels), or internalized. and selectivity through selective engineering of the
Immobilization of nanotransducers onto a substrate target neuron population. However, the viral vectors
or microelectrode facilitates the build-up of the used for the cell modification can cause immune
secondary signal to reach the threshold and control of responses, and the transfection is usually irreversible.
the nanotransducer distribution, but the implantation Thus, cell modification by genetic engineering causes
increases invasiveness. On the contrary, injectable safety and moral concerns and thwarts their clinical
and monodispersed nanotransducers minimize inva- applications.
siveness, but their distribution needs to be carefully Considering these critical factors, we herein review
controlled, as their placement and distribution affect four main types of nanotransducers for wireless neural
the cytotoxicity, modulation efficacy, efficiency, and stimulation: electrotransducers, thermotransducers,
consistency. For example, unbound and extracellularly optotransducers, and mechanotransducers. Each type
distributed nanotransducers prolong the stability of is further subdivided by the types of nanomaterial
the interface but tend to be removed by the fluid used. These nanotransducers are summarized regar-
circulation, such as the cerebrospinal fluid in the central ding these factors in Table S1 in the Electronic

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5082 Nano Res. 2018, 11(10): 5065–5106

Figure 6 Nanotransducer-assisted neural modulation. (a) Placement of nanotransducers: Nanotransducers can be dispersed
extracellularly (1), coated on a substrate (2), bound to the plasma membrane (3), targeted to receptors or anchoring molecules in the plasma
membrane (3’), targeted to ion channels in the plasma membrane (3”), and internalized (4). (b) and (c) Semiconducting nanorod-based
electrotransducers [170]. (b) Semiconducting nanorods convert light to electric fields. (c) A train of extracellular spikes triggered by a
light pulse. (d) and (e) Magnetoelectric nanomaterial-based electrotransducers [171]. (d) Magnetoelectric nanomaterials convert a
magnetic field to electric fields. (e) Modulated EEG waveforms. (f) and (g) Piezoelectric nanomaterial-based electrotransducers [172].
(f) Piezoelectric nanomaterials convert ultrasound to electric fields; (g) Ca2+ influxes following an ultrasound pulse. (h) and (i) Gold
nanomaterial-based thermotransducers [173]. (h) Gold nanomaterials convert light to heat. (i) A train of action potentials, each evoked by a
light pulse. (j) and (k) Superparamagnetic nanoparticle-based thermotransducers [163]. (j) Superparamagnetic nanoparticles convert a
magnetic field to heat. (k) Peristimulus time histogram. (l) and (m) Upconverting luminescent nanomaterial-based optotransducers
[174]. (l) Upconverting luminescent nanomaterials convert NIR light to visible light. (m) A train of action potentials, each evoked by an
NIR light pulse. (n) and (o) Superparamagnetic nanoparticle-based mechanotransducers [175]. (n) Superparamagnetic nanoparticles
convert a magnetic field to mechanical forces. (o) Peristimulus time histogram. (b) and (c) adapted with permission from Ref. [170], ©
the American Chemical Society 2014; (d) and (e) adapted with permission from Ref. [171], © Elsevier 2015; (f) and (g) adapted with
permission from Ref. [172], © the American Chemical Society 2015; (h) and (i) adapted with permission from Ref. [173], © Elsevier
2015; (j) and (k) adapted with permission from Ref. [163], © The American Association for the Advancement of Science 2015; (l) and
(m) adapted with permission from Ref. [174], © Royal Society of Chemistry 2015; (n) and (o) adapted with permission from Ref. [175],
© the American Chemical Society 2016.

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5083

Supplementary Material (ESM). Thus, efforts were mainly focused on improving


the interface stability by immobilizing QDs onto a
3.2.1 Nano electrotransducers
substrate.
Electrical signals are used to modulate neurons in the The initial contrivance of fNIs by immobilizing QDs
conventional modulation techniques. Instead of being onto salinized and poly-D-lysine-coated substrates
transmitted from an electrical stimulator to electrodes achieved slightly improved stability for 1 and 3 days,
through wires, electrical signals can also be converted respectively [182]. By culturing NG108-15 cells on a
locally by nanotransducers from other wirelessly layer-by-layer film containing repeating layers of HgTe
transmitted physical signals to target voltage-gated QDs and poly(dimethyldiallylammonium chloride),
ion channels for neural modulation [176, 177]. For Pappas et al. demonstrated that cells were depolarized
example, light, magnetic field, and ultrasound can be and even stimulated to fire action potentials upon
converted by QDs, magnetoelectric nanomaterials, or irradiation with 532 nm laser pulses (duration: 500 ms;
piezoelectric nanomaterials, respectively, into local irradiance: 800 mW/cm2) [185]. This was the first study
electric fields to modulate neurons. revealing an optical-to-electrical cellular modulation
QDs—From light to electric field: Upon excitation, assisted by QDs. Following the same strategy,
QDs convert light to an electric dipole (electron–hole CHO cells, RBL cells, NG108-15 cells, and primary
pair) via quantum confinement. The electric field hippocampal neurons were cultured on a stacked film
created by the electric dipole can serve as a secondary containing repeating layers of QDs and adhesives,
signal to modulate neurons [178]. The excitation and irradiation with 380 nm laser pulses depolarized
wavelengths of QDs range from ultraviolet via visible all types of cells and induced Ca2+ influxes and action
to near-infrared (NIR), and the tissue penetration potentials in NG108-15 cells and primary hippocampal
depth and spatial resolution of light depend on its neurons [186]. However, LnCap cells cultured on a
wavelength. The shorter the wavelength, the lower stacked film of CdTe QDs were hyperpolarized upon
the penetration and the higher resolution. NIR light illumination with 430 nm laser pulses [181]. In this
within the first (700–950 nm) and second (1,000– latter work, a simple coating strategy was used to
1,350 nm) NIR windows has minimal interactions drop-cast a film of CdSe QDs with van der Waals
(scattering and absorption) with tissues [179, 180], so force as the anchoring force. Cortical neurons cultured
this NIR light has maximal penetration through dermal on top were randomly hyperpolarized or depolarized
tissues, as deep as 4 mm, but the lowest resolution upon irradiation with 550 nm laser pulses, and
[164]. Thus, selection of QDs with their excitation cytotoxicity was found in these neurons. A CdSe
wavelengths depends on the target application. For QD-coated micropipette fell short to achieve consistent
example, light of short wavelengths is only suitable stimulation on cortical neurons, either.
for superficial neural modulation, such as in retinal It became clear that direct coating of QDs onto a
prostheses, while light of longer wavelengths can be substrate could not achieve a long-term stable fNI.
used for deeper intervention. As to the electric field, Although layer-by-layer QD films achieved some
mathematical modeling shows that an excited QD modulatory effects, their modulation efficiency and
could evoke an action potential if the distance between consistency were fairly low. In order to improve the
the QD and a voltage-gated ion channel is within stability of the fNI, instead of physically stacking layers
2 nm [181, 182]. Thus, intimate contacts of QDs to the of QDs, they were covalently bound to the substrate.
neuron’s membrane are realized by targeting the QDs Core–shell structured CdSe/CdS semiconducting
to the plasma membrane or by coating them onto a nanorods were bound to highly dense carbon nanotube
neuronal culturing substrate. films via carbodiimide crosslinking. Embryonic chick
However, due to their tiny size (2–6 nm), QDs are retinas, unresponsive to light, were cultured on the
easily internalized through endocytosis. Attempts to composite film and stimulated by 405 nm light pulses
create an fNI by targeting bio-conjugated QDs directly (duration: 100 ms; irradiance: 3–12 mW/cm2; Figs. 6(b)
to the plasma membrane rarely succeeded [182–184]. and 6(c)) [170]. The interface was stable for up to

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5084 Nano Res. 2018, 11(10): 5065–5106

21 days, with a significant improvement. However, transduction. Ultrasound, a longitudinal wave with a
due to the small photocurrents, the spatial resolution frequency higher than 20 Hz, can cause a nanoparticle
of this platform was low, with long latency and limited to vibrate along the transmission path and exert
synchronization. mechanical stress on it. Ultrasound can penetrate
Up to now, these QD-assisted wireless fNIs have deeply into soft tissues, and low-intensity, low-
only been studied in vitro, and the focus is still on frequency ultrasound can travel through the skull
establishing a stable interface to achieve effective and and be focused onto a specific deep brain region [188].
consistent modulation. The inherent tiny size of QDs Thus, ultrasound can serve as a primary wireless signal
makes it extremely challenging to use dispersed QDs. to activate piezoelectric nanotransducers to create a
One solution is to anchor QDs onto a substrate, and, secondary local electric field.
for this strategy, chemical bonding creates a more Zinc oxide nanowires, one of the most studied
stable interface than physical stacking or direct coating. piezoelectric nanomaterials, however, induce cyto-
Another solution is to explore semiconducting toxicity in excitable cells and fall short in maintaining
nanorods as an alternative. Even when the fNI is a stable interface in a physiological environment [189].
optimized with both chemical bonding and semicon- Other piezoelectric nanomaterials, including boron
ducting nanorods, its spatial resolution still needs nitride nanotubes [190, 191] and barium titanate
improvement. Additionally, the majority of QDs nanoparticles [172, 192], have been used to develop
are cytotoxic. Although surface modification and ultrasound-driven wireless cellular modulation
encapsulation have been used to alleviate this issue, techniques.
degradation of the QDs by the physiological acidic Similar to the development of QD-based fNIs, in
pH and enzymes is inevitable. the early efforts, dispersed piezoelectric boron nitride
Magnetoelectric nanomaterials—From magnetic field to nanotubes were studied. Dispersed piezoelectric
electric field: Magnetoelectric nanomaterials can convert boron nitride nanotubes were readily internalized
a magnetic field to an electric field via magnetoelectric by pheochromocytoma PC-12 cells, neuroblastoma
transduction. A computational study showed that SH-SY5Y cells, and myoblast C2C12 cells. When
such a transduction could theoretically recover the coupled with ultrasound, the piezoelectric boron nitride
deteriorated electric pulses in Parkinson’s disease [187]. nanotubes promoted neurite outgrowth of PC-12 cells
However, the model used a simplified assessment and SH-SY5Y cells as well as myogenesis of C2C12
of nanoparticles in an aqueous solution without cells [190, 191]. Dispersed gum Arabic-coated barium
considering many practical factors, such as the titanate nanoparticles, electrostatically attached to
physiological environment, placement of the nano- the plasma membrane of cell bodies and neurites of
particles, etc. A follow-up in vivo study demonstrated SH-SY5Y cells, induced high-amplitude Ca2+ influxes
that 30 nm CoFe2O4-BaTiO3 magnetoelectric core–shell when coupled with ultrasonic treatment (Figs. 6(f)
nanoparticles administrated into a mouse brain under and 6(g)) [172]. Following a similar placement strategy,
a low-energy magnetic field (100 Oersted; 0–20 Hz) dispersed gum Arabic-coated barium titanate nano-
could modulate the electroencephalography signals particles were electrostatically attached to the plasma
(Figs. 6(d) and 6(e)) [171]. However, further studies membrane of rat hippocampal neurons and cortical
are needed to assess the basics underlying this finding, neurons within neuronal networks cultured on a
such as the cellular distribution of the nanoparticles, microelectrode array [193]. Ultrasound treatment
biostability of the interface, spatial selectivity, temporal increased the spiking activity of these neuronal networks
precision, biochemistry, and, most importantly, the in a temporal pattern that matched the ultrasound
electrophysiological evidence of the cellular modulation pulses.
both in vitro and in vivo. Besides being applied in a dispersed form,
Piezoelectric nanomaterials—From ultrasound to electric piezoelectric nanomaterials were also immobilized onto
field: Piezoelectric nanomaterials convert mechanical substrates by blending piezoelectric nanoparticles
stress to an electric field via mechanoelectrical into polymers. A micropatterned photoresist blended

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5085

with barium titanate nanoparticles synergistically of the plasma membrane [198] and/or modulate
promoted the osteogenesis of osteosarcoma Saos-2 temperature-sensitive ion channels [199], therefore
cells under ultrasound [194]. With the same immo- modulating neural activities. Cell differentiation, such
bilization strategy, piezoelectric polyvinylidene fluoride as neurite outgrowth, has been used as an initial
trifluoroethylene (p(VDF-TrFE)) blended with barium hallmark to show the feasibility of this stimulation
titanate piezoelectric nanoparticles promoted neurite technique. More affirmative validations have also
outgrowth and induced Ca2+ transients in SH-SY5Y been provided involving in vitro and in vivo
cells cultured on both the p(VDF-TrFE) film and the electrophysiological tests.
blended film under ultrasound [192]. However, both Internalized gold nanorods coupled with NIR light
films decreased the cellular proliferation compared treatment were initially shown to promote neuronal
to a control Ibidi film. differentiation. Gold nanorods (resonant wavelength:
It is worth noting that also piezoelectric polymers 780 nm) were internalized in NG108-15 neuronal cells.
coupled with ultrasound were explored for neural Neurite outgrowth was promoted upon irradiation
modulation. A piezoelectric p(VDF-TrFE) film coupled with a 780 nm continuous-wave laser (duration: 1 min;
with ultrasound promoted the neurogenesis of PC-12 irradiance: 0–13.75 W/cm2) [200], and intracellular
cells [195]. Moreover, electrospun piezoelectric p(VDF- Ca2+ transients were induced upon irradiation with
TrFE) microfibers alone promoted neurite extension 780 nm laser pulses (duration: 20–100 ms; radiant
of dorsal root ganglion neurons [196]. However, no exposure: 0.07–407 J/cm2) [167].
synergetic treatment of piezoelectric polymer nanofibers Untargeted dispersed gold nanorods were found to
and ultrasound has been reported yet. stimulate neurons with NIR illumination proportionally
Although theoretically piezoelectric nanomaterials to the laser duration or irradiance. In vitro rat auditory
can convert mechanical stimuli into an electric field, neurons were treated with silica-coated gold nanorods
no direct evidence has validated the generation of an (resonant wavelength: 780 nm), and the nanorods
electric field from dispersed piezoelectric nanomaterials were distributed extracellularly, contacted the plasma
under ultrasound. Many questions still remain to be membrane, or were internalized [201]. A laser pulse
answered: How does ultrasound interact with the of 780 nm (duration: 0.025–50 ms; power: 90 mW;
dispersed nanoparticles? Can adequate electric fields irradiance: unknown) induced inward transmembrane
be generated? How do the generated electric fields currents and membrane depolarization proportional
distribute? Additionally, the long-term interface stability to the pulse duration. However, the neural responses
and biocompatibility of dispersed piezoelectric nano- were not consistent among the tested neurons, and a
materials must also be addressed. train of action potentials matching the pulse pattern
was not achieved. Gold nanorods (resonant wavelength:
3.2.2 Nano thermotransducers
977 nm) were injected into a rat sciatic nerve in vivo
Gold nanomaterials—From light to heat: Rod- and and extracellularly distributed [202]. Irradiation
sphere-shaped gold nanoparticles can convert NIR with a 980-nm laser pulse (duration: 1 ms; irradiance:
and visible light to localized heat through surface 159–1,046 W/cm2) enhanced the amplitude of compound
plasmon resonance, respectively [197]. Like QDs, the nerve action potentials proportionally to the radiant
excitation (resonant) wavelength of gold nanomaterials exposure. Comparing to a null control without gold
affects the penetration depth and spatial resolution, nanorods, irradiation with gold nanorods increased
and its selection depends on the target application. the responsivity by five times and substantially
The resonant wavelength can be tuned by controlling lowered the threshold by two-thirds.
the aspect ratio [197]. Incident irradiation with the To improve the stimulation specificity and efficiency,
corresponding resonant wavelength excites confined gold nanomaterials were targeted to the plasma
electron oscillation and collision at the nanoparticle membrane and ion channels, and illumination induced
surface, resulting in heat generation and dissipation. excitatory or inhibitory effects in stimulated cells.
The generated local heat can change the capacitance Amine-terminated PEGylated gold nanorods (resonant

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5086 Nano Res. 2018, 11(10): 5065–5106

wavelength: 785 nm) were bound to the plasma this work, the ligand-conjugated gold nanoparticles
membrane of rat hippocampal, cortical, and olfactory were also injected into acute mouse hippocampal
bulb neurons, and irradiation of 785 nm laser pulses slices ex vivo, and irradiation with 532 nm laser
(duration: 10 s; irradiance: 1.5 W/cm2) inhibited the pulses (duration: 10 ms; power: 140 mW) induced a
spontaneous, electrically stimulated, and chemically membrane depolarization indicated by a VSD.
hyperexcited neural network activity, while the unbound To improve the interface stability and stimulation
nanorods did not substantially affect the neural network efficiency, gold nanomaterials were immobilized
spike rate [203]. Cationized high-density lipoprotein- onto substrates. A monolayer of amine-terminated
coated gold nanorods (resonant wavelength: 785 nm) PEGylated gold nanorods was coated onto a micro-
were targeted to the plasma membrane of HEK293T electrode array for rat hippocampal neuronal culturing,
human epithelial cells, and irradiation with a 780 nm and irradiation with 785 nm laser pulses (duration:
continuous-wave laser (duration: 120 s; irradiance: 120 s; irradiance: 0.3–2.1 W/cm2) inhibited the neural
800 W/cm2) induced Ca2+ influxes [169]. Uncoated gold activities and blocked signal transmission between
nanoparticles (resonant wavelength: 565 nm) and neural clusters through axons [206]. The authors
antibody-conjugated PEGylated gold nanoparticles claimed that the nature of the effects depended on the
(resonant wavelength: 568 nm) were bound to the radiant exposure, i.e., a higher radiant exposure excited
plasma membrane of non-genetically engineered the neurons, whereas a lower radiant exposure inhibited
and genetically engineered rat hippocampal neurons, them. In another study, a gold nanoparticle (resonant
respectively, and irradiation with 800 nm laser pulses wavelength: 532 nm)-coated glass micropipette was
(duration: 140 fs; irradiance: 0.27–1.02 mW/cm2) placed in the vicinity of a neuroblastoma SH-SY5Y
induced Ca2+ transients at both the cellular and cell or a rat cardiomyocyte, and the cellular activity
subcellular level and evoked action potentials [204]. could be stimulated (duration: 1–5 ms; power: 75–
Streptavidin-coated gold nanorods (resonant wavelength: 120 mW) or inhibited (duration: 300 ms; power:
977 nm) were targeted to the biotinylated antibodies 120 mW) upon irradiation with 532 nm laser pulses
bound to the plasma membrane of rat hippocampal [207]. The authors claimed that the excitatory or
neurons, and irradiation with a 980 nm laser pulse inhibitory effect depended on the pulse duration, i.e., a
(duration: 400 μs; irradiance: 75.25 W/cm2) improved shorter pulse of a few milliseconds excited the neurons,
the stimulation efficiency, lowered the threshold, and whereas a longer pulse of tens of milliseconds to
shortened the latency in comparison to the null minutes inhibited the neurons. However, based on a
control without gold nanorods [205]. In this study, complete analysis of this topic (Table 1), neither
antibody-conjugated gold nanorods were also injected assumption could fully account for all the observations.
into the rat motor cortex, and irradiation with a 980-nm To discern the excitatory and inhibitory effects, the
laser pulse (duration: 1.5 ms; irradiance: 85.3 W/cm2) consideration may not be solely on the primary
induced whisker oscillations. Following the same signal—the light, but should also be given to many
targeting strategy, streptavidin-coated gold nanorods other factors, particularly the heat generated locally.
(resonant wavelength: 982 nm) were also targeted Heat is the secondary signal with which the neurons
to the biotinylated antibodies bound to the plasma directly interact. Both the amount and rate of heat
membrane of rat astrocytes, and irradiation with generation affect the neural responses, which are
a 980 nm laser pulse (duration: 950 μs; irradiance: determined by the light (duration and irradiance),
1,381 W/cm2) induced intracellular Ca2+ transients nanoparticle amount and placement, etc.
[162]. Ligand-conjugated gold nanoparticles (resonant Moreover, gold nanomaterials also have a great
wavelength: 523 nm) were bound to the ion channels tendency to be rapidly internalized, making it difficult
in the plasma membrane of rat dorsal root ganglion to maintain a stable and effective fNI. The uptake of
neurons in vitro, and irradiation with 532 nm laser gold nanomaterials can also cause inconsistency and
pulses (duration: 1 ms; irradiance: 31 kW/cm2) evoked cytotoxicity [167, 201]. Furthermore, although NIR
a train of action potentials (Figs. 6(h) and 6(i)) [173]. In light can penetrate deeper than visible light, their

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5087

Table 1 Summary of gold nanomaterial-assisted thermotransduction


Reference Laser duration (ms) Laser irradiance (W/cm2) Irradiant exposure (mJ/cm2) Modulation effect
a
[201] 0.025–50 N/A N/A Excitatory
[202] 1 159–1,046 159–1,046 Excitatory
5 5
[203] 6 × 10 1.5 9 × 10 Inhibitory
6 9
[169] 7.2 × 10 800 5.76 × 10 Excitatory
–10 –4 –3 –14 –13
[204] 1.4 × 10 2.7×10 –1.02×10 3.78×10 –1.428×10 Excitatory
0.4 75.25 30.1 Excitatory
[205]
1.5 85.3 127.95 Excitatory
[162] 0.95 1,381 1,311.95 Excitatory
[173] 1 31,000 31,000 Excitatory
6 6 7
[206] 7.2 × 10 0.3–2.1 2.16×10 –1.512×10 Inhibitory
1–5 75–120 75–600 Excitatory
[207]
300 N/A N/A Inhibitory
a
N/A: no available information.

penetration depth is still limited to 4 mm in dermal particles were targeted to genetically engineered
tissues [164]. Additionally, heat, as a localized secondary protein markers in the plasma membrane of HEK293T
signal, can be limited by its temporal responsivity human epithelial cells and rat hippocampal neurons
[167, 173, 201]. The heat generation must be carefully expressing temperature-gated transient receptor
controlled, as overheating can lead to protein potential vanilloid 1 (TRPV1) ion channels. Application
denaturation and cell damage [208]. of a radio-frequency (RF) magnetic field generated
Nevertheless, gold nanomaterials have a good highly localized heat, inducing Ca2+ influxes in
biocompatibility, and their surface plasmonic resonant HEK293T cells and evoking action potentials in
properties can be tailored based on the application hippocampal neurons within 45 s [11]. In this study,
needs. Gold nanorods also have a higher heat the PEG-phospholipid-coated nanoparticles were also
generation efficiency, thus requiring a nanoparticle injected into the amphids of C. elegans having TRPV1
concentration almost 1,000 times less than the magnetic ion channel-expressing sensory neurons, and appli-
nanoparticle-assisted thermotransduction discussed cation of a magnetic field initiated the worm’s
below [169]. More importantly, they do not require thermal avoidance response within 5 s. Besides being
genetic engineering of the target cells. Thus, gold anchored onto the plasma membrane, antibody-
nanomaterials have become one of the most developed conjugated magnetic nanoparticles were also directly
types of wireless nanotransducers. targeted to the TRPV1 ion channels in the plasma
Superparamagnetic nanomaterials—From magnetic field membrane of HEK293T cells and mouse embryonic
to heat: Superparamagnetic nanoparticles generate stem cells expressing a Ca2+-dependent promoter for
localized heat when exposed to an oscillating or insulin secretion, and application of an RF magnetic
alternating magnetic field. Similar to the gold field resulted in cytosolic Ca2+ increase, enhanced
nanomaterial-assisted thermotransduction, this localized proinsulin release, and relevant gene expressions [209].
heating phenomenon via hysteresis triggered by In this study, antibody-conjugated magnetic nano-
a magnetic field makes these nanoparticles good particles were also injected into xenografts of PC-12
candidates as wireless nanotransducers for cellular pheochromocytoma cells expressing both TRPV1 ion
modulation. This type of technique started with the channels and Ca2+-dependent promoter for insulin
targeted placement of these nanoparticles. secretion in mouse, and application of a magnetic field
Streptavidin-conjugated superparamagnetic nano- resulted in a decrease of blood glucose, an increase

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5088 Nano Res. 2018, 11(10): 5065–5106

of plasma insulin, and an increase of relevant gene naturally occurring. In this study, superparamagnetic
expressions. Additionally, genetically engineered cells nanoparticles were also injected into the transfected
expressing ferritin fusion protein, TRPV1 ion channels, ventral tegmental area of the mouse brain, and
and a Ca2+-dependent promoter for insulin secretion application of a magnetic field increased the c-fos
produced magnetic ferritin nanoparticles intracellularly; gene expression. These nanotransducers lasted more
the application of a magnetic field also resulted in than a month in vivo.
increases of proinsulin release and relevant gene For magnetic nanoparticle-assisted thermotrans-
expressions. In this publication, the Ca2+ influxes were duction, heat was confirmed as the localized secondary
induced faster than in Ref. [11], and it was the first signal. However, mechanical force as another possible
one to use endogenous nanoparticles for cellular secondary signal cannot be excluded [210]. Similar
modulation. to gold nanomaterial-based thermotransducers, the
Stanley et al. extended the use of endogenous ferritin temporal resolution of this type of techniques also
nanoparticles for wireless cellular modulation in needs to be significantly improved. Interestingly, the
another study [210]. In genetically engineered HEK293T thermotransduction enabled by magnetic nanoparticles
human epithelial cells and mouse mesenchymal generally requires genetic engineering to overexpress
stem cells expressing ferritin proteins (cytoplasmic, TRPV1 ion channels and other binding proteins, while
membrane-bound, or TRPV1-bound), TRPV1 ion gold nanomaterial-assisted thermotransduction usually
channels, and a Ca2+-dependent promoter for insulin does not involve genetic engineering. Although some
secretion, the intracellularly produced ferritin nano- of the differences can be attributed to the difference
particles coupled with an RF magnetic field resulted in experimental designs, the exemption of genetic
in increases in proinsulin release and relevant gene engineering may have taken advantage of the higher
expressions. In this publication, in vivo studies were heating efficiency of gold nanomaterials. To generate
conducted by either implanting gelatin scaffolds con- the same amount of heat, a much larger number of
taining the genetically engineered mouse mesenchymal magnetic nanoparticles is needed compared to gold
stem cells in mouse or injecting adenoviruses to nanomaterials [169]. Nevertheless, superparamagnetic
transfect cells. The application of an RF magnetic field nanoparticles have good biocompatibility and
resulted in a decrease of blood glucose, an increase of biodegradability. And the primary signal, an RF
plasma insulin, and an increase of relevant gene magnetic field, can penetrate more deeply than light
expressions. The effectiveness was shown to last for at into tissues without causing tissue heating. Additionally,
least several weeks. Application of a static magnetic magnetic nanoparticles can also convert a magnetic
field led to similar results in vitro and in vivo only for field into mechanical forces, a secondary signal that
TRPV1-bound ferritin proteins, indicating that both can also interact with cells, which will be discussed
heating and mechanical force could activate TRPV1 in the mechanotransducers below.
ion channels.
3.2.3 Nano optotransducers
Untargeted placement of superparamagnetic nano-
particles was also implemented for cellular modulation Upconverting luminescent nanomaterials—From NIR light
to suppress internalization and prolong the interface. to visible light: Optogenetics (to be discussed in detail
PEGylated poly(acrylic acid)-coated superparamagnetic below), artificially inserting light-sensitive ion channels
nanoparticles were distributed in the adjacent into the plasma membrane and modulating neural
surroundings of HEK-293FT cells and hippocampal activity with light, has become a powerful tool for
neurons, and application of an alternating magnetic neural stimulation with its exceptional spatiotemporal
field induced Ca2+ influxes in HEK293-FT cells and resolution and specificity [211]. The majority of light-
evoked trains of action potentials in hippocampal sensitive ion channels, such as the channelrhodopsins
neurons within 5 s (Figs. 6(j) and 6(k)) [163]. It is worth (ChRs), respond to blue to green light with some
mentioning that the cells were genetically engineered newly developed ChRs responding to orange to red
to express higher levels of TRPV1 ion channels than light, such as the red-activatable channelrhodopsin

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5089

(ReaChR) [212], Chrimson [213], and Jaws [214] with 980 nm continuous-wave NIR light induced Ca2+
through laborious engineering [174]. However, visible influxes in HeLa cells, activated the Ca2+/NFAT pathway
light only penetrates soft tissues superficially. Thus, in T lymphocytes, and promoted melanoma tumor
optogenetics often demands the implantation of a light destruction by promoting dendritic cell maturation
source close to the target neuron population. On the and the sensitivity of T lymphocytes to antigen.
other hand, NIR light (650–1,450 nm) can penetrate Immobilizing core–shell structured upconverting
deeper into the soft tissues when the wavelength luminescent nanomaterials onto substrates improved
lies within the biological window [215]. Additionally, the modulation efficiency and resolution. The ChR2
upconverting luminescent nanomaterials can be (maxima absorbance: 470–475 nm)-expressing mouse
excited by NIR light and emit ultraviolet/visible light hippocampal neurons were cultured on a poly(lactic-
through an upconversion process. These inspire the co-glycolic acid) scaffold containing NaYF4:Yb/
development of a new wireless neural modulation Tm@NaYF4 core–shell structured nanoparticles (excitation:
technique by combining upconverting luminescent 980 nm; maxima emission: 475 nm), and irradiation
nanomaterials with optogenetics. The upconverting with a sequence of 980 nm NIR laser pulses induced
luminescent nanomaterial-assisted optogenetics, in a train of action potentials (Figs. 6(l) and 6(m)) [174].
comparison to the original optogenetics, uses an More importantly, in this study, a single pulse could
inexpensive light source and could possibly avoid rapidly evoke a single action potential. Similarly,
surgical implantation of the light source. The emission ChR-expressing rat hippocampal neurons were cultured
wavelength of such optotransducers can be relatively on a cover glass over a poly(methyl methacrylate) film
easily tailored instead of a laborious screening and containing IR-806-sensitized NaYF4:Yb/Er@NaYF4/Yb
engineering of the ion channels [174]. core–shell structured nanoparticles (excitation: 800 nm;
Dispersed untargeted and targeted placement of maxima emission: ~ 540 nm), and irradiation with a
the upconverting luminescent nanomaterials could sequence of 800 nm continuous-wave NIR light pulses
be coupled with optogenetics to modulate cells. evoked light intensity-dependent depolarization and
Synergistic treatment by NaYF4:Sc/Yb/Er nanoparticles a train of action potentials, the temporal pattern
(excitation: 980 nm; maxima emission: 543 nm) and a of which matched that of the light pulses [219]. The
976 nm NIR laser pulse induced photocurrents in improved efficiency and resolution could be due to
C1V1 (maxima absorbance: 539 nm) expressing mouse the enhanced upconversion efficiency of core–shell
neuroblastoma and rat neuron hybrid ND7/23 cells structured nanoparticles and increased interface
[216]. However, the cellular distribution of the nano- stability.
particles was not assessed in this study, and a single It is worth noting that although the above excitation
light pulse induced a series of action potentials due wavelength (975–980 nm) is within the NIR range, its
to the slow off-kinetics of C1V1 channels. Quasi- absorption by water is still significant [220]. It is also
continuous-wave NIR light (980 nm) was used to worth mentioning that NIR light can only penetrate
excite extracellularly and intracellularly distributed through dermal tissues up to 4 mm [164] and has
silica-coated NaYF4:Yb/Tm nanoparticles (excitation: difficulties to penetrate through bones and the skull.
980 nm; maxima emission: 450 nm), and the emission Additionally, for these upconverting luminescent
induced Ca2+ influxes in ChR2-expressing HEK293T nanomaterial-assisted optogenetic techniques, both
cells and elicited reversal responses in C. elegans the optotransducers and ion channels need to be well
having ChR2-expressing mechanosensory neurons designed. On the one hand, a high upconversion
[217]. Core–shell structured NaYF4:Yb/Tm@NaYF4 efficiency is greatly demanded to use a low-power
nanoparticles (excitation: 980 nm; maxima emission: NIR source to reach the light threshold for ion channel
470 nm) were made to improve the upconversion activation while avoiding thermal hazards to the cells
efficiency, and the nanoparticles were conjugated with [217]. The emission of ultraviolet light and its toxicity
streptavidin and bound to the Ca2+ channels engineered also need to be assessed [217]. In addition, in vivo
with a streptavidin-binding site [218]. Illumination biocompatibility of the upconverting luminescent

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5090 Nano Res. 2018, 11(10): 5065–5106

nanomaterials needs to be further improved and channels and induced almost instantaneous Ca2+
optimized for clinical stimulation [217]. On the other fluxes [230].
hand, the ion channels should have rapid dynamics Magnetic nanoparticles bound to the cell surface
and stable expression. receptors were magnetized and aggregated upon
exposure to a magnetic field. The receptor clustering
3.2.4 Nano mechanotransducers
due to particle aggregation activated receptor
Superparamagnetic nanomaterials—From magnetic field to oligomerization-dependent cell signaling pathways.
mechanical force: Comparing to other recently developed Ligand-conjugated superparamagnetic nanoparticles
nanotransducer-assisted cellular modulation techniques, were targeted to the transmembrane FcεRI receptors
mechanical manipulation of cellular responses via of RBL-2H3 rat mast cells, and application of a
magnetic nanoparticles has been studied much earlier. magnetic field activated the Ca2+ signaling pathway,
Magnetic nanoparticles were bound to the cell surface, increasing the cytosolic Ca2+ concentration [231].
and application of a magnetic field generated either Similarly, antibody-conjugated magnetic nanoparticles
piconewton forces to stretch the corresponding ion were targeted to death receptors at the surface of
channels and cytoskeleton via particle twisting or DLD-1 human colon cancer cells and injected into
pulling [221, 222] or femtonewton forces to induce zebrafish embryos, and magnetic field-induced receptor
receptor clustering via particle aggregation. Inter- clustering activated an apoptosis signaling pathway
nalized magnetic nanoparticles were also used to [232]. Antibody-conjugated or streptavidin-functionalized
manipulate protein distributions inside the cells [223, superparamagnetic nanoparticles were targeted to the
224]. Interestingly, while the majority of magnetic epidermal growth factor receptors (EGFRs) of A431
nanoparticle-assisted mechanical stimulations were human epidermoid carcinoma cells and genetically-
conducted on non-excitable cells, the concepts can be engineered chimeric EGFRs of HeLa cells, respectively,
transferred to neural modulation straightforwardly. and magnetic field-induced receptor clustering activated
Magnetic nanoparticles can be bound to the cell the EGFR signaling pathway [233].
surface, and through twisting or pulling of these Internalized magnetic nanoparticles could be con-
nanoparticles, a magnetic field can stretch the ion trolled by an external magnetic field to manipulate
channels and the cytoskeleton. For example, endothelial the intracellular protein and organelle distribution.
cells were bound to ferromagnetic microparticles via The superparamagnetic nanoparticles internalized
a molecular linker, and shear stress was applied into cortical neurons coupled with a magnetic field
to the integrin receptors via the molecular linker by exerted magnetic forces to modulate protein tau
magnetic twisting, leading to cytoskeleton stiffening distribution and cell polarization [223]. The super-
[225–227]. Collagen-coated magnetic microbeads were paramagnetic nanoparticles were microinjected into
bound to human gingival fibroblasts, and the HeLa cells, and a magnetic field was applied to
application of magnetic force on the plasma membrane control the distribution of HaloTag proteins [224].
promoted Ca2+ influxes via mechanosensitive channels Similarly, anti-TrKB agonist antibody-functionalized
[228]. Antibody- or complex-conjugated magnetic superparamagnetic nanoparticles were rapidly
nanoparticles were targeted to a genetically added internalized into TrKB signaling endosomes of primary
loop region of the mechanosensitive TREK-1 ion neurons and promoted neurite outgrowth [234]. A
channels in the plasma membrane of COS-7 kidney focal magnetic force halted the neurite outgrowth via
cells, and the application of a magnetic field induced modulation of the location of these TrKB signaling
an increase of the outward current [229]. Similarly, endosomes.
magnetic microparticles were bound to the integrin Magnetic nanoparticles were recently employed to
receptors and chimeric integrin receptors of bovine modulate neurons. Starch-coated ferromagnetic nano-
capillary endothelial cells, and the magnetic pulling on particles were mainly bound to the plasma membrane
these integrin receptors activated the mechanosensitive of cortical neurons, and chitosan-coated ferromagnetic
transient receptor potential vanilloid 4 (TRPV4) ion nanoparticles were mostly internalized (Figs. 6(n) and

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5091

6(o)) [175]. Application of a high gradient magnetic in biological systems. Once stimulated by extrinsic
field for both placements of nanoparticles induced signals or stimuli, spatiotemporal changes within
Ca2+ fluxes in the neural network via activation of neurons will lead to profound effects on behavior.
N-type mechanosensitive ion channels. This mechanism Optogenetics offers many advantages, such as cell-type
was further validated in a follow-up study [208]. In specific targeting, high spatiotemporal resolution
this work, starch-coated magnetic nanoparticles were (subcellular and millisecond), and concurrent recording
bound to the plasma membrane of the neurons in with metal electrodes without the introduction of
which the N-type mechanosensitive ion channels were aberrant artifacts. For a comprehensive overview, the
over-expressed, and chronic treatment with a high readers are referred to many excellent reviews,
gradient magnetic field lowered the expression of these including Refs. [1, 236–240].
N-type mechanosensitive ion channels by chronically In 2005, Boyden et al. reported that mammalian
activating these channels. Further electrophysiological neurons could be precisely modulated by light with
validations are needed for these wireless mechano- the introduction of a single microbial ChR gene [211].
transducers to be used for neural modulation, and Opsins are seven-transmembrane domain proteins
their responsivity and temporal resolution need to be (structurally akin to G-protein coupled receptors)
improved. possessing the chromophore retinal, which undergoes
Considering the widespread existence of mechano- a conformational change in response to photon
sensitive ion channels and cell signaling pathways absorption and drives ion transport across the cell’s
[222], these mechanical cellular modulations on non- membrane. Commonly, bacteriorhodopsin (BR), ChR,
excitable cells can be adapted and tailored for neural and halorhodopsin (HR) are the most widely used
modulation. Magnetic microparticles are limited for opsin families that can modulate neural activity in
in vivo cellular modulation due to their relatively response to different wavelengths of light (Fig. 7(a)).
large size [11]. Magnetic nanoparticles are usually Two different effects exhibited by these proteins
bio-friendly and can be rapidly cleared in the human are excitation and inhibition of neural activity. BRs
body, thus, these nanoparticles are used clinically in are green-light-activated ion pumps that transport
magnetic resonance imaging. Magnetic nanoparticles hydrogen ions out the cytoplasm causing hyper-
bound to receptors could also be rapidly internalized polarization and preventing an action potential. ChRs
[232, 233]. Their biodegradability and tendency for are blue-light-activated ion channels that conduct
internalization make it difficult to maintain a stable various cations (Na+, H+, K+, and Ca2+) across the
fNI lasting for a long time [235]. Similar to the wireless cellular membrane and depolarize neurons, eliciting
thermotransducers based on magnetic nanoparticles, an action potential. HRs are yellow-light-activated ion
it is necessary to discern if the secondary signal is pumps that transport Cl– ions intracellularly and
heat or mechanical force so that the design can be inhibit neuron firing [241].
optimized. Introduction of opsin genes is accomplished
via transgenic (Cre-LoxP) or viral delivery (adeno-
3.3 Bionanotransducer-enabled functional neural associated virus (AAV), Lentivirus) methods. Cre
stimulation transgenic mice can be crossed with another
mouse strain expressing the opsin of interest in a
3.3.1 Optogenetics
Cre-dependent manner, an example of which was
Nanobiotechnology utilizing light-sensitive proteins accomplished with Ai39 mice with Cre-dependent
can function as a switch to control neural activities like Halo3.0/eNpHR3.0 expression [242]. Viral gene delivery
intracellular signal transduction pathways, action using an AAV is another method where a plasmid
potential propagation, and even gene transcription containing an opsin gene under the control of a
or protein–protein interactions. Optogenetics is the promoter (CMV) would be injected into a brain region
combination of genetic and optical methods to allow or transfected into cells in vitro and visualized with a
for targeted, fast control of precisely defined events fluorescent reporter gene [243]. Owing to the specificity

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5092 Nano Res. 2018, 11(10): 5065–5106

Figure 7 Optogenetic molecular applications. (a) Main opsin families that conduct ions upon light stimulation: BR (H+), HR (Cl–),
and ChR (Na+, K+, Ca2+, H+) activated via green, yellow, and blue lights, respectively [244]. (b) Schematic representation of ReaChR
consisting of the N-terminal domain from ChEF, the transmembrane domains from VChR1 except that domain F is from VChR2 and an
L171I mutation in domain C [212]. (c) and (d) Protein control via light stimulation [245]. (c) Transmembrane protein control over
intracellular signaling cascade upon light stimulation. (d) Caged ligands released upon light stimulation causing subsequent protein
conformational change and leading to activation. (e) Spectra of emission and activation wavelengths of luciferase and eNpHR3.0 with
peaks at ~ 550 and ~ 600 nm, respectively [246]. (f) and (g) Photoactivatable CRISPR-Cas9 system [247, 248]. (f) Cleaved Cas9 protein
whose photoinducible domains (pMag/nMag) dimerize upon blue-light stimulation, allowing for Cas9 fragments to reassociate and cut
DNA at a specific region. (g) Dead-Cas9 tagged with CIB1, targeted to a promoter region, serves as a transcription modulator upon
blue-light stimulation causing CRY2 tagged with a gene activator to bind to CIB1 and promote transcription. (a) adapted with
permission from Ref. [244], © Elsevier 2011; (b) adapted with permission from Ref. [212], © Springer Nature 2013; (c) and (d) adapted
with permission from Ref. [245], © Springer Nature 2014; (e) adapted with permission from Ref. [246], © Frontiers Media S. A. 2014;
(f) and (g) adapted with permission from Refs. [247, 248], © Springer Nature 2015 and Elsevier 2015, respectively.

of optogenetics, selection of similar promoter sequences tissue penetration with blue light having the shallowest
to target a cell subtype will allow a cell-specific penetration depth (1–2 mm) and near-infrared light
integration since cells that are expressing the same having the deepest (4 mm). Aside from inserting an
promoters will have the required transcription factors optic fiber deep into the brain or using upconverting
to initiate opsin gene synthesis. Transgenic mice luminescent optotransducers as discussed above,
offer a robust and precise optogenetic activation or several opsins have been discovered and engineered
suppression in vivo. While viral delivery methods have to have shifted light sensitivities and faster ion
a few limitations, including limited DNA packaging transport to compensate for this deficiency. An
space and transfection efficiency, it still remains a inhibitory HR from Natronomonas pharaonis (Halo3.0/
viable option to genetically modify specific cells. eNpHR3.0) was developed, which possessed a
A distinct limitation of optical stimulation is its membrane localization tag that could hyperpolarize

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5093

neurons upon yellow-light stimulation with increased (Fig. 7(e)). Endogenous light production from those
excitability and faster photokinetics [249–252]. VChR1, cells would then stimulate nearby cells expressing an
a ChR from Volvox carteri, can be activated at 589 nm, opsin and further control cell signaling and intracellular
which is slightly red-shifted when compared to ChR2. events [246].
However, it expressed poorly in mammalian cells and Recent advances in gene editing and targeting
had minimal trafficking to the cellular membrane methods have supplemented the field of optogenetics
[253]. VChR1 was further modified with components with the advent of photoactivatable CRISPR-Cas9.
of other ChR proteins and an L171I mutation One method split Cas9 into two domains, each tagged
allowing for its optimal excitation to be red-shifted with a photoinducible domain that, when exposed
further (630 nm); termed ReaChR, this opsin had to light, dimerized and brought the Cas9 domains
vastly improved membrane trafficking and kinetic together allowing for it to target and edit the genome
properties (Fig. 7(b)) [212]. Engineered for prolonged based upon the guide RNA (gRNA) (Fig. 7(f)) [247].
periods of neural depolarization, bistable step-function The same group also demonstrated earlier the ability
opsins (SFOs) allowed for orders of magnitude greater to use dead-Cas9 (dCas9) as a photoactivatable
light sensitivity and had lasting effects without the transcription system. dCas9 is a useful tool for gene
need for continuous light stimulation [254]. Persistent regulation since it can still target DNA with gRNA,
work on opsins for H+/Cl– conducting-based inhibition but lacks nuclease activity, and can be coupled to gene
is ongoing, for they could mediate the optical silencing activators and repressors. Tagged with CIB1, dCas9
of significant neural activities over relevant timescales induced gene transcription once CRY2, tagged with an
[249, 255–258]. Opsins with faster kinetics that could activator, heterodimerized with CIB1 upon blue-light
deliver more efficient photocurrents while also being stimulation (Fig. 7(g)) [248]. Utilizing one or adopting
expressed at higher levels within the cell are of great multiple methods concerning optogenetics can allow
paramount [259, 260]. for a high degree of specificity and spatiotemporal
Aside from using light to stimulate an opsin control over neural stimulation.
within a cell to drive a response, another way is to
3.3.2 Magneto(thermo)genetics
use chemically modified or genetically encoded caged
ligands to control neural activities (Figs. 7(c) and 7(d)) Due to its noninvasiveness and deep tissue penetration,
[261, 262]. Light-sensitive proteins, engineered with magnetic neural stimulation is of great interest. To
the attachment of photo-switchable tethered ligands, stimulate neurons with cell-type specificity, magnetic
could ultimately regulate the interaction between nanoparticles combined with genetic engineering of
ligand and receptor and allow for precise control over certain ion channels have been developed for nano-
intracellular signaling events. The correct wavelength of transducer-assisted magnetic stimulation [263, 264],
light would effectively activate the ligand, transitioning as discussed earlier. The commonly used channels
from a caged to an uncaged state, allowing for a include mechanosensitive channels (TREK) [175, 229]
protein–protein interaction to occur. However, several and heat-sensitive capsaicin receptor channels (TRP)
considerations have limited the application of caged [11, 163]. TREK-1 is a type of tandem-pore K +
proteins, one being that cleaving the cage tends to be channel (2PK+) and produces an outwardly rectifying
irreversible and may produce toxic byproducts that K+ leak current to regulate the resting membrane
can alter cell signaling. Reversible caged ligands, potential and cellular excitability [265]. Hughes et al.
using a short connecting molecule called azobenzene, inserted a 6-histidine-repeat into the TREK-1 channel
controls the activity of ligands with the ability to and attached an anti-His antibody-grafted magnetic
go through many photoactivation cycles without a nanoparticle to it to directly activate the mechano-
reduction in activity [245, 261]. An alternative approach, sensitive channel by exerting a mechanical force
typically used concurrently with other optogenetic through a permanent magnetic field (Fig. 8(a)) [229].
methods, uses constitutively expressed luciferin (a Wheeler et al. developed the Magneto, a new fusion
light producing compound) within a specific cell type of the nonselective cation channel TRPV4 and the

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5094 Nano Res. 2018, 11(10): 5065–5106

paramagnetic protein ferritin (an iron-containing


storage protein), and applied it to mediate Ca2+ influx
upon magnetic stimulation to modulate the behavior
of freely moving zebrafish and mice [266]. TRP
channels form the most important temperature-
sensitive ion channel family, among which TRPM8
and TRPA1 are sensitive to cold temperatures, and
TRPV1, TRPV3, and TRPV4 respond to an elevation
in local temperature [267]. For example, TPRV1 opens
when the local temperature increases from normal
to 42 °C and causes an influx of Ca2+ to depolarize
the cell. These ion channels can be heated by an
RF magnetic field through nano thermotransducers
(e.g., superparamagnetic nanoparticles) or bionano
thermotransducers (e.g., ferritin) either attaching
to a ligand inserted into the membrane (Fig. 8(b))
or directly attaching to the ion channel through an
engineered ligand recognition (Fig. 8(c)) [11, 209, 210].
The latter approach of specific ion-channel binding
ensures a higher efficiency in signal transduction
while minimizing heat-induced side-effects. In their
study, Stanley et al. used genetically encoded ferritin
as the thermotransducers to control insulin expression
and blood glucose in mice [210]. The ferritin was
targeted to TPRV1 for gating under low RF waves, Figure 8 Mechanisms of magneto(thermo)genetics and sono-
which resulted in elevated intracellular Ca2+ concen- genetics. (a) Magnetomechanical stimulation with mechanotransducers
trations and induced the expression of proinsulin. to modulate the mechanosensitive channel (TREK1). His-tagged
magnetic nanoparticles are targeted to TREK1 with a 6-histidine
They also expanded this ferritin-TPRV1 magneto-
repeat inserted region [229]. (b) Magnetothermal stimulation with
thermogenetics system to target hypothalamic different nanoparticle placements. Top, streptavidin-conjugated
glucose-sensing neurons to control proinsulin release nanoparticles bound to the plasma membrane around TRPV1
in mice [268]. channels through a genetically engineered linker [11]. Bottom,
Additionally, Qin et al. reported a single magnetic through His tag binding, superparamagnetic nanoparticles (grey
circle) or ferritin proteins (purple circle) are attached to the
protein MagR for neural stimulation [268]. Long et al.
thermal-sensitive channel (TPR family) [209, 210]. (c) Potential
invented a noninvasive magnetogenetics that combined mechanism for sonogenetics, involving the mechanosensitive TRP-4
the genetic targeting of a magnetoreceptor with channel [271]. (a) adapted with permission from Ref. [229], © The
remote magnetic stimulation through the use of a Royal Society 2008; (b) top adapted with permission from
Ref. [11], © Springer Nature 2010; (b) bottom adapted with
genetically encoded magnetic protein [270], without
permission from Refs. [209, 210], © The American Association
the requirement for injection of exogenous super- for the Advancement of Science 2012 and Springer Nature 2015,
paramagnetic nanoparticles. Although these emerging respectively; (c) adapted with permission from Ref. [271],
techniques have aroused a great deal of enthusiasm in © Springer Nature 2015.
the field, there are still many technical obstacles, and
3.3.3 Sonogenetics
further efforts are required to decipher the specific
stimulation mechanism and develop new magnetic- Ultrasound can also be used to noninvasively stimulate
sensitive receptors. the nervous system. To enhance the power efficiency

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5095

and target specificity, genetically encoded acoustic- mode (“1” for action potential and “0” for no action
sensitive ion channels can be used as the bionano- potential) by each electrode, with a great loss of the
transducers. An early investigation by Ibsen et al. used information contents that can otherwise be conveyed
the misexpressed ultrasonic-sensitive channel TRP-4, through the fNI. While some electrode types exist to
a pore-forming subunit of a mechanotransduction record the subthreshold neuronal activities [273],
channel, and ultrasound-amplified microbubbles to they are, at present, limited to cell culture interfacing
manipulate the functions of interneurons and sensory only. The need to develop implantable versions of
neurons in C. elegans for behavior control (Fig. 8(c)) such “analog” neural interfaces is urgent.
[271]. They termed their technology “sonogenetics”. Furthermore, the plastic nature of neuron–neuron
Kubanek et al. measured the effects of ultrasound communication has never been purposely emulated
on mechanosensitive ion channels embedded within at any fNI. We argue that the goal of neuroelectrode
cellular membranes in the Xenopus oocyte system, design in the functional aspect should not be set to
including potassium mechanosensitive channels achieve a static signal-transmission interface but rather
(TREK-1, TREK-2, TRAAK) and a sodium-sensitive to achieve a dynamic interface that remodels based
channel (Nav1.5) [272]. Ultrasound could modulate on a memory of the signals transmitted by it. These
the currents flowing through the ion channels by gaps in interface efficiency and plasticity, thus, call
up to 23%. Compared to other bionanotransducer- for an urgent and serious rethinking of the design of
enabled neural stimulation techniques, sonogenetics fNIs to capture the full vocabulary of a neuron and
has a much longer way to go. Much more work is still implement plasticity into the interface. With such
needed to elucidate the stimulation mechanism, search new fNIs, it would be more effective and efficient to
for more effective acoustic-sensitive ion channels that wire a neuromorphic chip running biologically realistic
can be transfected into other cells, and expend the
neuronal models to the brain to repair and even
application to mammalian neuron stimulation.
augment the brain [274], as exemplified in the recent
attempts in building cognitive prostheses [275, 276].
4 Future perspectives
The three thematic nano fNIs covered in this review, Acknowledgements
i.e., nanoelectrode-based, nanotransducer-assisted,
L. G. is supported by The Defense Advanced Research
and bionanotransducer-enabled, represent the most
Projects Agency (No. D17AP00031) of the USA. The
exciting new advances in the field of neural interface
views, opinions, and/or findings contained in this
engineering over the past 15 years, which address the
article are those of the author and should not be
long-term physical integration issue of the neural
interpreted as representing the official views or policies,
interface. These fascinating advances embody the
either expressed or implied, of the Defense Advanced
insensibility and indistinguishability principles that
Research Projects Agency or the Department of
we proposed earlier [2] through the pursuits of
Defense.
miniaturization and biomimicry and the employment
C. X. is supported by National Institute of
of nature’s materials and designs.
Tackling the inefficiency problem of limited com- Neurological Disorders and Stroke through Grant #
munication bandwidth through the abiotic–biotic R01NS102917, Welch Foundation through Grant #
interface has so far solely relied on increasing the F-1941-20170325, and Department of Defense through
electrode’s spatial resolution, density, and number. Clinical and Rehabilitative Medicine Research Program
However, it has been largely ignored that, in the current under award No. W81XWH-16-1-0580.
design of neuroelectrodes, the neuron’s “vocabulary”
(including action potentials, excitatory and inhibitory Electronic Supplementary Material: Supplementary
synaptic potentials [273]) in forming the brain’s material (a summary table on nanotransducer-assisted
language is “filtered” into an overly simplistic binary wireless neural modulation) is available in the online

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5096 Nano Res. 2018, 11(10): 5065–5106

version of this article at https://doi.org/10.1007/ injury induced by electrical-stimulation. IEEE Trans. Biomed.
s12274-018-2127-4. Eng. 1990, 37, 996–1001.
[13] Kim, J. H.; Manuelidis, E. E.; Glen, W. W.; Kaneyuki, T.
Diaphragm pacing: Histopathological changes in the phrenic
References nerve following long-term electrical stimulation. J. Thorac.
Cardiovasc. Surg. 1976, 72, 602–608.
[1] Fenno, L.; Yizhar, O.; Deisseroth, K. The development and
[14] Kotov, N. A.; Winter, J. O.; Clements, I. P.; Jan, E.; Timko,
application of optogenetics. Ann. Rev. Neurosci. 2011, 34,
B. P.; Campidelli, S.; Pathak, S.; Mazzatenta, A.; Lieber, C.
389–412.
M.; Prato, M. et al. Nanomaterials for neural interfaces.
[2] Guo, L. The pursuit of chronically reliable neural interfaces:
Adv. Mater. 2009, 21, 3970–4004.
A materials perspective. Front. Neurosci. 2016, 10, 599.
[15] Wang, Y. C.; Guo, L. Nanomaterial-enabled neural stimulation.
[3] Marin, C.; Fernández, E. Biocompatibility of intracortical
Front. Neurosci. 2016, 10, 69.
microelectrodes: Current status and future prospects. Front.
[16] Young, A. T.; Cornwell, N.; Daniele, M. A. Neuro-nano
Neuroeng. 2010, 3, 8.
interfaces: Utilizing nano-coatings and nanoparticles to
[4] Polikov, V. S.; Tresco, P. A.; Reichert, W. M. Response of
enable next-generation electrophysiological recording, neural
brain tissue to chronically implanted neural electrodes. J.
stimulation, and biochemical modulation. Adv. Funct.
Neurosci. Methods 2005, 148, 1–18.
Mater. 2017, 28, 1700239.
[5] Barrese, J. C.; Rao, N.; Paroo, K.; Triebwasser, C.; Vargas-
[17] Krack, P.; Batir, A.; Van Blercom, N.; Chabardes, S.;
Irwin, C.; Franquemont, L.; Donoghue, J. P. Failure mode
Fraix, V.; Ardouin, C.; Koudsie, A.; Limousin, P. D.;
analysis of silicon-based intracortical microelectrode arrays
Benazzouz, A.; LeBas, J. F. et al. Five-year follow-up of
in non-human primates. J. Neural Eng. 2013, 10, 066014.
bilateral stimulation of the subthalamic nucleus in advanced
[6] Kozai, T. D. Y.; Catt, K.; Li, X.; Gugel, Z. V.; Olafsson,
Parkinson’s disease. N. Engl. J. Med. 2003, 349, 1925–1934.
V. T.; Vazquez, A. L.; Cui, X. T. Mechanical failure modes
[18] Boon, P.; Raedt, R.; De Herdt, V.; Wyckhuys, T.; Vonck,
of chronically implanted planar silicon-based neural probes
K. Electrical stimulation for the treatment of epilepsy.
for laminar recording. Biomaterials 2015, 37, 25–39.
Neurotherapeutics 2009, 6, 218–227.
[7] McConnell, G. C.; Rees, H. D.; Levey, A. I.; Gutekunst,
[19] Burgess, N. The 2014 Nobel Prize in physiology or medicine:
C.-A.; Gross, R. E.; Bellamkonda, R. V. Implanted neural
A spatial model for cognitive neuroscience. Neuron 2014,
electrodes cause chronic, local inflammation that is correlated
84, 1120–1125.
with local neurodegeneration. J. Neural Eng. 2009, 6, 056003.
[20] Kim, T. I.; McCall, J. G.; Jung, Y. H.; Huang, X.; Siuda,
[8] Kozai, T. D. Y.; Langhals, N. B.; Patel, P. R.; Deng, X. P.;
E. R.; Li, Y.; Song, J.; Song, Y. M.; Pao, H. A.; Kim, R. H.
Zhang, H. N.; Smith, K. L.; Lahann, J.; Kotov, N. A.; Kipke,
et al. Injectable, cellular-scale optoelectronics with applications
D. R. Ultrasmall implantable composite microelectrodes
for wireless optogenetics. Science 2013, 340, 211–216.
with bioactive surfaces for chronic neural interfaces. Nat.
[21] Kozai, T. D. Y.; Du, Z. H.; Gugel, Z. V.; Smith, M. A.;
Mater. 2012, 11, 1065–1073.
Chase, S. M.; Bodily, L. M.; Caparosa, E. M.; Friedlander,
[9] Xie, C.; Liu, J.; Fu, T. M.; Dai, X. C.; Zhou, W.; Lieber,
R. M.; Cui, X. T. Comprehensive chronic laminar single-unit,
C. M. Three-dimensional macroporous nanoelectronic
multi-unit, and local field potential recording performance
networks as minimally invasive brain probes. Nat. Mater.
with planar single shank electrode arrays. J. Neurosci.
2015, 14, 1286–1292.
Methods 2015, 242, 15–40.
[10] Luan, L.; Wei, X. L.; Zhao, Z. T.; Siegel, J. J.; Potnis, O.;
[22] Fraser, G. W.; Schwartz, A. B. Recording from the same
Tuppen, C. A.; Lin, S. Q.; Kazmi, S.; Fowler, R. A.;
neurons chronically in motor cortex. J. Neurophysiol. 2012,
Holloway, S. et al. Ultraflexible nanoelectronic probes form
107, 1970–1978.
reliable, glial scar–free neural integration. Sci. Adv. 2017, 3, [23] Perge, J. A.; Homer, M. L.; Malik, W. Q.; Cash, S.;
e1601966. Eskandar, E.; Friehs, G.; Donoghue, J. P.; Hochberg, L. R.
[11] Huang, H.; Delikanli, S.; Zeng, H.; Ferkey, D. M.; Pralle, Intra-day signal instabilities affect decoding performance in
A. Remote control of ion channels and neurons through an intracortical neural interface system. J. Neural. Eng.
magnetic-field heating of nanoparticles. Nat. Nanotechnol. 2013, 10, 036004.
2010, 5, 602–606. [24] Gilletti, A.; Muthuswamy, J. Brain micromotion around
[12] McCreery, D. B.; Agnew, W. F.; Yuen, T. G. H.; Bullara, L. implants in the rodent somatosensory cortex. J. Neural. Eng.
Charge-density and charge per phase as cofactors in neural 2006, 3, 189–195.

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5097

[25] Prasad, A.; Xue, Q. S.; Dieme, R.; Sankar, V.; Mayrand, R. Wilkinson, C. D. An extracellular microelectrode array for
C.; Nishida, T.; Streit, W. J.; Sanchez, J. C. Abiotic-biotic monitoring electrogenic cells in culture. Biosens. Bioelectron.
characterization of Pt/Ir microelectrode arrays in chronic 1990, 5, 223–234.
implants. Front. Neuroeng. 2014, 7, 2. [38] Pine, J. Recording action potentials from cultured neurons
[26] Prasad, A.; Xue, Q. S.; Sankar, V.; Nishida, T.; Shaw, G.; with extracellular microcircuit electrodes. J. Neurosci.
Streit, W. J.; Sanchez, J. C. Comprehensive characterization Methods 1980, 2, 19–31.
and failure modes of tungsten microwire arrays in chronic [39] Hai, A.; Spira, M. E. On-chip electroporation, membrane
neural implants. J. Neural. Eng. 2012, 9, 056015. repair dynamics and transient in-cell recordings by arrays of
[27] Gilgunn, P. J.; Ong, X. C.; Flesher, S. N.; Schwartz, A. B.; gold mushroom-shaped microelectrodes. Lab Chip 2012, 12,
Gaunt, R. A. Structural analysis of explanted microelectrode 2865–2873.
arrays. In Proceedings of the 2013 6th International [40] Xie, C.; Lin, Z. L.; Hanson, L.; Cui, Y.; Cui, B. X. Intracellular
IEEE/EMBS Conference on Neural Engineering (NER), San recording of action potentials by nanopillar electroporation.
Diego, CA, USA, 2013, pp 719–722. Nat. Nanotechnol. 2012, 7, 185–190.
[28] Patel, P. R.; Na, K.; Zhang, H. N; Kozai, T. D. Y.; Kotov, [41] Xie, C.; Hanson, L.; Xie, W. J.; Lin, Z. L.; Cui, B. X.;
N. A.; Yoon, E.; Chestek, C. A. Insertion of linear 8.4 μm Cui, Y. Noninvasive neuron pinning with nanopillar arrays.
diameter 16 channel carbon fiber electrode arrays for single Nano Lett. 2010, 10, 4020–4024.
unit recordings. J. Neural Eng. 2015, 12, 046009. [42] Lin, Z. C.; Xie, C.; Osakada, Y.; Cui, Y.; Cui, B. X. Iridium
[29] Rousche, P. J.; Normann, R. A. Chronic recording capability oxide nanotube electrodes for sensitive and prolonged
of the Utah Intracortical Electrode Array in cat sensory intracellular measurement of action potentials. Nat. Commun.
cortex. J. Neurosci. Methods 1998, 82, 1–15. 2014, 5, 3206.
[30] Williams, J. C.; Rennaker, R. L.; Kipke, D. R. Long-term [43] Robinson, J. T.; Jorgolli, M.; Shalek, A. K.; Yoon, M.-H.;
neural recording characteristics of wire microelectrode arrays Gertner, R. S.; Park, H. Vertical nanowire electrode arrays as
implanted in cerebral cortex. Brain Res. Protoc. 1999, 4, a scalable platform for intracellular interfacing to neuronal
303–313. circuits. Nat. Nanotechnol. 2012, 7, 180–184.
[31] Kipke, D. R.; Vetter, R. J.; Williams, J. C.; Hetke, J. F. [44] Abbott, J.; Ye, T. Y.; Qin, L.; Jorgolli, M.; Gertner, R. S.;
Silicon-substrate intracortical microelectrode arrays for Ham, D.; Park, H. CMOS nanoelectrode array for all-
long-term recording of neuronal spike activity in cerebral electrical intracellular electrophysiological imaging. Nat.
cortex. IEEE Trans. Neural Syst. Rehabil. Eng. 2003, 11, Nanotechnol. 2017, 12, 460–466.
151–155. [45] Liu, R.; Chen, R. J.; Elthakeb, A. T.; Lee, S. H.; Hinckley,
[32] Simeral, J. D.; Kim, S. P.; Black, M. J.; Donoghue, J. P.; S.; Khraiche, M. L.; Scott, J.; Pre, D.; Hwang, Y.; Tanaka,
Hochberg, L. R. Neural control of cursor trajectory and A. et al. High density individually addressable nanowire
click by a human with tetraplegia 1000 days after implant arrays record intracellular activity from primary rodent and
of an intracortical microelectrode array. J. Neural Eng. human stem cell derived neurons. Nano Lett. 2017, 17,
2011, 8, 025027. 2757–2764.
[33] Sakmann, B.; Neher, E. Single-Channel Recording, 2nd ed.; [46] Tian, B. Z.; Cohen-Karni, T.; Qing, Q.; Duan, X. J.; Xie, P.;
Springer: New York, NY, USA, 2009. Lieber, C. M. Three-dimensional, flexible nanoscale field-
[34] Souslova, V.; Cesare, P.; Ding, Y. N.; Akopian, A. N.; effect transistors as localized bioprobes. Science 2010, 329,
Stanfa, L.; Suzuki, R.; Carpenter, K.; Dickenson, A.; 830–834.
Boyce, S.; Hill, R. et al. Warm-coding deficits and aberrant [47] Qing, Q.; Jiang, Z.; Xu, L.; Gao, R. X.; Mai, L. Q.; Lieber,
inflammatory pain in mice lacking P2X3 receptors. Nature C. M. Free-standing kinked nanowire transistor probes for
2000, 407, 1015–1017. targeted intracellular recording in three dimensions. Nat.
[35] Lee, J.; Ishihara, A.; Oxford, G.; Johnson, B.; Jacobson, K. Nanotechnol. 2014, 9, 142–147.
Regulation of cell movement is mediated by stretch- [48] Duan, X. J.; Gao, R. X.; Xie, P.; Cohen-Karni, T.; Qing, Q.;
activated calcium channels. Nature 1999, 400, 382–386. Choe, H. S.; Tian, B. Z.; Jiang, X. C.; Lieber, C. M.
[36] Thomas, C. A., Jr.; Springer, P. A.; Loeb, G. E.; Berwald- Intracellular recordings of action potentials by an extracellular
Netter, Y.; Okun, L. M. A miniature microelectrode array to nanoscale field-effect transistor. Nat. Nanotechnol. 2012, 7,
monitor the bioelectric activity of cultured cells. Exp. Cell 174–179.
Res. 1972, 74, 61–66. [49] Fu, T.-M.; Duan, X. J.; Jiang, Z.; Dai, X. C.; Xie, P.; Cheng,
[37] Connolly, P.; Clark, P.; Curtis, A. S.; Dow, J. A. T.; Z. G.; Lieber, C. M. Sub-10-nm intracellular bioelectronic

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5098 Nano Res. 2018, 11(10): 5065–5106

probes from nanowire–nanotube heterostructures. Proc. Natl. electrodes for extracellular recording from electrogenic
Acad. Sci. USA 2014, 111, 1259–1264. cells. Nanotechnology 2011, 22, 265104.
[50] Jayant, K.; Hirtz, J. J.; Plante, I. J.-L.; Tsai, D. M.; de Boer, [62] Zhou, H. B.; Li, G.; Sun, X. N.; Zhu, Z. H.; Jin, Q. H.; Zhao,
W. D. A. M.; Semonche, A.; Peterka, D. S.; Owen, J. S.; J. L.; Ren, Q. S. Integration of Au nanorods with flexible
Sahin, O.; Shepard, K. L. et al. Targeted intracellular voltage thin-film microelectrode arrays for improved neural
recordings from dendritic spines using quantum-dot-coated interfaces. J. Microelectromech. Syst. 2009, 18, 88–96.
nanopipettes. Nat. Nanotechnol. 2017, 12, 335–342. [63] Zhao, Z. Y.; Gong, R. X.; Zheng, L.; Wang, J. In vivo neural
[51] Kim, R.; Nam, Y. Electrochemical layer-by-layer approach recording and electrochemical performance of microelectrode
to fabricate mechanically stable platinum black microelectrodes arrays modified by rough-surfaced AuPt alloy nanoparticles
using a mussel-inspired polydopamine adhesive. J. Neural with nanoporosity. Sensors 2016, 16, 1851.
Eng. 2015, 12, 026010. [64] Kim, Y. H.; Kim, G. H.; Kim, M. S.; Jung, S.-D. Iridium
[52] Li, M. Z.; Zhou, Q.; Duan, Y. Y. Nanostructured porous oxide–electrodeposited nanoporous gold multielectrode array
platinum electrodes for the development of low-cost fully with enhanced stimulus efficacy. Nano Lett. 2016, 16,
implantable cortical electrical stimulator. Sensor. Actuat. B: 7163–7168.
Chem. 2015, 221, 179–186. [65] Zeng, Q.; Xia, K.; Sun, B.; Yin, Y. L.; Wu, T. Z.; Humayun,
[53] Weremfo, A.; Carter, P.; Hibbert, D. B.; Zhao, C. M. S. Electrodeposited iridium oxide on platinum nanocones
Investigating the interfacial properties of electrochemically for improving neural stimulation microelectrodes. Electrochim.
roughened platinum electrodes for neural stimulation. Acta 2017, 237, 152–159.
Langmuir 2015, 31, 2593–2599. [66] Jan, E.; Hendricks, J. L.; Husaini, V.; Richardson-Burns, S.
[54] Park, S.; Song, Y. J.; Boo, H.; Chung, T. D. Nanoporous Pt M.; Sereno, A.; Martin, D. C.; Kotov, N. A. Layered carbon
microelectrode for neural stimulation and recording: In vitro nanotube-polyelectrolyte electrodes outperform traditional
characterization. J. Phys. Chem. C 2010, 114, 8721–8726. neural interface materials. Nano Lett. 2009, 9, 4012–4018.
[55] Lee, Y. J.; Lee, S. J.; Yoon, H. S.; Park, J. Y. A bulk [67] Deng, M.; Yang, X.; Silke, M.; Qiu, W. M.; Xu, M. S.;
micromachined silicon neural probe with nanoporous platinum Borghs, G.; Chen, H. Z. Electrochemical deposition of
electrode for low impedance recording. In SENSORS, 2013 polypyrrole/graphene oxide composite on microelectrodes
IEEE, Baltimore, MD, USA, 2013, pp 1–4. towards tuning the electrochemical properties of neural
[56] Chung, T.; Wang, J. Q.; Wang, J.; Cao, B.; Li, Y.; Pang, S. probes. Sensor. Actuat. B: Chem. 2011, 158, 176–184.
W. Electrode modifications to lower electrode impedance [68] Luo, X. L.; Weaver, C. L.; Tan, S. S.; Cui, X. T. Pure
and improve neural signal recording sensitivity. J. Neural graphene oxide doped conducting polymer nanocomposite
Eng. 2015, 12, 056018. for bio-interfacing. J. Mater. Chem. B 2013, 1, 1340–1348.
[57] Chen, Y.-C.; Hsu, H.-L.; Lee, Y.-T.; Su, H.-C.; Yen, S.-J.; [69] Weaver, C. L.; Li, H.; Luo, X.; Cui, X. T. A graphene
Chen, C.-H.; Hsu, W.-L.; Yew, T.-R.; Yeh, S.-R.; Yao, D.-J. oxide/conducting polymer nanocomposite for electrochemical
et al. An active, flexible carbon nanotube microelectrode dopamine detection: Origin of improved sensitivity and
array for recording electrocorticograms. J. Neural Eng. 2011, specificity. J. Mater. Chem. B 2014, 2, 5209–5219.
8, 034001. [70] Ng, A. M. H.; Kenry; Teck Lim, C.; Low, H. Y.; Loh, K. P.
[58] Kim, G. H.; Kim, K.; Nam, H.; Shin, K.; Choi, W.; Shin, Highly sensitive reduced graphene oxide microelectrode
J. H.; Lim, G. CNT-Au nanocomposite deposition on gold array sensor. Biosens. Bioelectron. 2015, 65, 265–273.
microelectrodes for improved neural recordings. Sensor. [71] Kook, G.; Lee, S. W.; Lee, H. C.; Cho, I.-J.; Lee, H. J. Neural
Actuat. B: Chem. 2017, 252, 152–158. probes for chronic applications. Micromachines 2016, 7, 179.
[59] Kim, J.-H.; Kang, G.; Nam, Y.; Choi, Y.-K. Surface- [72] Jorfi, M.; Skousen, J. L.; Weder, C.; Capadona, J. R.
modified microelectrode array with flake nanostructure for Progress towards biocompatible intracortical microelectrodes
neural recording and stimulation. Nanotechnology 2010, 21, for neural interfacing applications. J. Neural Eng. 2014, 12,
085303. 011001.
[60] Kim, Y. H.; Kim, G. H.; Kim, A. Y.; Han, Y. H.; Chung, [73] Kozai, T. D. Y.; Jaquins-Gerstl, A. S.; Vazquez, A. L.;
M.-A.; Jung, S.-D. In vitro extracellular recording and Michael, A. C.; Cui, X. T. Brain tissue responses to neural
stimulation performance of nanoporous gold-modified multi- implants impact signal sensitivity and intervention strategies.
electrode arrays. J. Neural Eng. 2015, 12, 066029. ACS Chem. Neurosci. 2015, 6, 48–67.
[61] Brüggemann, D.; Wolfrum, B.; Maybeck, V.; Mourzina, Y.; [74] Takmakov, P.; Ruda, K.; Scott Phillips, K.; Isayeva, I. S.;
Jansen, M.; Offenhäusser, A. Nanostructured gold micro- Krauthamer, V.; Welle, C. G. Rapid evaluation of the

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5099

durability of cortical neural implants using accelerated Acad. Sci. USA 2013, 110, 6694–6699.
aging with reactive oxygen species. J. Neural Eng. 2015, [86] Liu, J.; Fu, T. M.; Cheng, Z. G.; Hong, G. S.; Zhou, T.; Jin,
12, 026003. L. H.; Duvvuri, M.; Jiang, Z.; Kruskal, P.; Xie, C. et al.
[75] Seymour, J. P.; Kipke, D. R. Neural probe design for Syringe-injectable electronics. Nat. Nanotechnol. 2015, 10,
reduced tissue encapsulation in CNS. Biomaterials 2007, 28, 629–636.
3594–3607. [87] Fu, T.-M.; Hong, G. S.; Zhou, T.; Schuhmann, T. G.;
[76] Skousen, J. L.; Merriam, M. E.; Srivannavit, O.; Perlin, G.; Viveros, R. D.; Lieber, C. M. Stable long-term chronic
Wise, K. D.; Tresco, P. A. Reducing surface area while brain mapping at the single-neuron level. Nat. Methods
maintaining implant penetrating profile lowers the brain 2016, 13, 875–882.
foreign body response to chronically implanted planar [88] Luan, L.; Sullender, C. T.; Li, X.; Zhao, Z. T.; Zhu, H. L.;
silicon microelectrode arrays. Prog. Brain Res. 2011, 194, Wei, X. L.; Xie, C.; Dunn, A. K. Nanoelectronics enabled
167–180. chronic multimodal neural platform in a mouse ischemic
[77] Karumbaiah, L.; Saxena, T.; Carlson, D.; Patil, K.; Patkar, model. J. Neurosci. Methods 2018, 295, 68–76.
R.; Gaupp, E. A.; Betancur, M.; Stanley, G. B.; Carin, L.; [89] Zhang, H. N.; Patel, P. R.; Xie, Z. X.; Swanson, S. D.;
Bellamkonda, R. V. Relationship between intracortical Wang, X. D.; Kotov, N. A. Tissue-compliant neural
electrode design and chronic recording function. Biomaterials implants from microfabricated carbon nanotube multilayer
2013, 34, 8061–8074. composite. ACS Nano 2013, 7, 7619–7629.
[78] Guitchounts, G.; Markowitz, J. E.; Liberti, W. A.; Gardner, [90] Henze, D. A.; Borhegyi, Z.; Csicsvari, J.; Mamiya, A.;
T. J. A carbon-fiber electrode array for long-term neural Harris, K. D.; Buzsáki, G. Intracellular features predicted
recording. J. Neural Eng. 2013, 10, 046016. by extracellular recordings in the hippocampus in vivo. J.
[79] Vitale, F.; Summerson, S. R.; Aazhang, B.; Kemere, C.; Neurophysiol. 2000, 84, 390–400.
Pasquali, M. Neural stimulation and recording with bidirectional, [91] Du, J. G.; Blanche, T. J.; Harrison, R. R.; Lester, H. A.;
soft carbon nanotube fiber microelectrodes. ACS Nano Masmanidis, S. C. Multiplexed, high density electrophysiology
2015, 9, 4465–4474. with nanofabricated neural probes. PLoS One 2011, 6,
[80] Mercanzini, A.; Cheung, K.; Buhl, D. L.; Boers, M.; e26204.
Maillard, A.; Colin, P.; Bensadoun, J.-C.; Bertsch, A.; Renaud, [92] Marblestone, A. H.; Zamft, B. M.; Maguire, Y. G.; Shapiro,
P. Demonstration of cortical recording using novel flexible M. G.; Cybulski, T. R.; Glaser, J. I.; Amodei, D.; Stranges,
polymer neural probes. Sensor. Actuat. A: Phys. 2008, 143, P. B.; Kalhor, R.; Dalrymple, D. A. et al. Physical principles
90–96. for scalable neural recording. Front. Comput. Neurosci.
[81] Wu, F.; Tien, L. W.; Chen, F.; Berke, J. D.; Kaplan, D. L.; 2013, 7, 137.
Yoon, E. Silk-backed structural optimization of high-density [93] Camuñas-Mesa, L. A.; Quiroga, R. Q. A detailed and fast
flexible intracortical neural probes. J. Microelectromech. model of extracellular recordings. Neural Comput. 2013, 25,
Syst. 2015, 24, 62–69. 1191–1212.
[82] Du, Z. J.; Kolarcik, C. L.; Kozai, T. D. Y.; Luebben, S. D.; [94] Pedreira, C.; Martinez, J.; Ison, M. J.; Quiroga, R. Q. How
Sapp, S. A.; Zheng, X. S.; Nabity, J. A.; Cui, X. T. Ultrasoft many neurons can we see with current spike sorting
microwire neural electrodes improve chronic tissue integration. algorithms? J. Neurosci. Methods 2012, 211, 58–65.
Acta Biomater. 2017, 53, 46–58. [95] Guo, L.; DeWeerth, S. P. An effective lift-off method for
[83] Sohal, H. S.; Clowry, G. J.; Jackson, A.; O’Neill, A.; Baker, patterning high-density gold interconnects on an elastomeric
S. N. Mechanical flexibility reduces the foreign body substrate. Small 2010, 6, 2847–2852.
response to long-term implanted microelectrodes in rabbit [96] Khodagholy, D.; Doublet, T.; Quilichini, P.; Gurfinkel, M.;
cortex. PLoS One 2016, 11, e0165606. Leleux, P.; Ghestem, A.; Ismailova, E.; Hervé, T.; Sanaur,
[84] Guo, L.; Guvanasen, G. S.; Liu, X.; Tuthill, C.; Nichols, T. S.; Bernard, C. et al. In vivo recordings of brain activity
R.; DeWeerth, S. P. A PDMS-based integrated stretchable using organic transistors. Nat. Commun. 2013, 4, 1575.
microelectrode array (isMEA) for neural and muscular [97] Viventi, J.; Kim, D.-H.; Vigeland, L.; Frechette, E. S.;
surface interfacing. IEEE Trans. Biomed. Circuits Syst. Blanco, J. A.; Kim, Y.-S.; Avrin, A. E.; Tiruvadi, V. R.;
2013, 7, 1–10. Hwang, S.-W.; Vanleer, A. C. et al. Flexible, foldable,
[85] Liu, J.; Xie, C.; Dai, X. H.; Jin, L. H.; Zhou, W.; Lieber, actively multiplexed, high-density electrode array for
C. M. Multifunctional three-dimensional macroporous mapping brain activity in vivo. Nat. Neurosci. 2011, 14,
nanoelectronic networks for smart materials. Proc. Natl. 1599–1605.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5100 Nano Res. 2018, 11(10): 5065–5106

[98] Rios, G.; Lubenov, E. V.; Chi, D.; Roukes, M. L.; Siapas, A. [109] Grienberger, C.; Konnerth, A. Imaging calcium in neurons.
G. Nanofabricated neural probes for dense 3-D recordings of Neuron 2012, 73, 862–885.
brain activity. Nano Lett. 2016, 16, 6857–6862. [110] Peterka, D. S.; Takahashi, H.; Yuste, R. Imaging voltage
[99] Scholvin, J.; Kinney, J. P.; Bernstein, J. G.; Moore- in neurons. Neuron 2011, 69, 9–21.
Kochlacs, C.; Kopell, N.; Fonstad, C. G.; Boyden, E. S. [111] Rowland, C. E.; Susumu, K.; Stewart, M. H.; Oh, E.;
Close-packed silicon microelectrodes for scalable spatially Mäkinen, A. J.; O’Shaughnessy, T. J.; Kushto, G.; Wolak,
oversampled neural recording. IEEE Trans. Biomed. Eng. M. A.; Erickson, J. S.; Efros, A. L. et al. Electric field
2016, 63, 120–130. modulation of semiconductor quantum dot photoluminescence:
[100] Wu, F.; Stark, E.; Ku, P.-C.; Wise, K. D.; Buzsáki, G.; Insights into the design of robust voltage-sensitive cellular
Yoon, E. Monolithically integrated μLEDs on silicon neural imaging probes. Nano Lett. 2015, 15, 6848–6854.
probes for high-resolution optogenetic studies in behaving [112] Alivisatos, A. P. Semiconductor clusters, nanocrystals,
animals. Neuron 2015, 88, 1136–1148. and quantum dots. Science 1996, 271, 933–937.
[101] Wei, X. L.; Luan, L.; Zhao, Z. T.; Li, X.; Zhu, H. L.; [113] Empedocles, S. A.; Bawendi, M. G. Quantum-confined
Potnis, O.; Xie, C. Nanofabricated ultraflexible electrode stark effect in single CdSe nanocrystallite quantum dots.
arrays for high-density intracortical recording. Adv. Sci. Science 1997, 278, 2114–2117.
2018, 1700625. [114] Marshall, J. D.; Schnitzer, M. J. Optical strategies for
[102] Lopez, C. M.; Andrei, A.; Mitra, S.; Welkenhuysen, M.; sensing neuronal voltage using quantum dots and other
Eberle, W.; Bartic, C.; Puers, R.; Yazicioglu, R. F.; Gielen, semiconductor nanocrystals. ACS Nano 2013, 7, 4601–4609.
G. G. E. An implantable 455-active-electrode 52-channel [115] Park, K.; Weiss, S. Design rules for membrane-embedded
CMOS neural probe. IEEE J. Solid-St. Circ. 2014, 49, voltage-sensing nanoparticles. Biophys. J. 2017, 112,
248–261. 703–713.
[103] Lopez, C. M.; Putzeys, J.; Raducanu, B. C.; Ballini, M.; [116] Knöpfel, T.; Díez-García, J.; Akemann, W. Optical probing
Wang, S. W.; Andrei, A.; Rochus, V.; Vandebriel, R.; of neuronal circuit dynamics: Genetically encoded versus
Severi, S.; Hoof, C. V. et al. A neural probe with up to 966 classical fluorescent sensors. Trends Neurosci. 2006, 29,
electrodes and up to 384 configurable channels in 0.13 μm 160–166.
SOI CMOS. IEEE Trans. Biomed. Circuits Syst. 2017, 11, [117] Tsytsarev, V.; Liao, L. D.; Kong, K. V.; Liu, Y. H.;
510–522. Erzurumlu, R. S.; Olivo, M.; Thakor, N. V. Recent progress
[104] Jun, J. J.; Steinmetz, N. A.; Siegle, J. H.; Denman, D. J.; in voltage-sensitive dye imaging for neuroscience. J.
Bauza, M.; Barbarits, B.; Lee, A. K.; Anastassiou, C. A.; Nanosci. Nanotechnol. 2014, 14, 4733–4744.
Andrei, A.; Aydın, Ç. et al. Fully integrated silicon probes [118] Tsytsarev, V.; Pope, D.; Pumbo, E.; Yablonskii, A.;
for high-density recording of neural activity. Nature 2017, Hofmann, M. Study of the cortical representation of
551, 232–236. whisker directional deflection using voltage-sensitive dye
[105] Kruss, S.; Salem, D. P.; Vuković, L.; Lima, B.; Vander optical imaging. NeuroImage 2010, 53, 233–238.
Ende, E.; Boyden, E. S.; Strano, M. S. High-resolution [119] Eriksson, D.; Wunderle, T.; Schmidt, K. Visual cortex
imaging of cellular dopamine efflux using a fluorescent combines a stimulus and an error-like signal with a
nanosensor array. Proc. Natl. Acad. Sci. USA 2017, 114, proportion that is dependent on time, space, and stimulus
1789–1794. contrast. Front. Syst. Neurosci. 2012, 6, 26.
[106] Beyene, A. G.; McFarlane, I. R.; Pinals, R. L.; Landry, M. [120] Grinvald, A.; Salzberg, B. M.; Cohen, L. B. Simultaneous
P. Stochastic simulation of dopamine neuromodulation for recording from several neurones in an invertebrate central
implementation of fluorescent neurochemical probes in nervous system. Nature 1977, 268, 140–142.
the striatal extracellular space. ACS Chem. Neurosci. [121] Cohen, L. B.; Salzberg, B. M.; Grinvald, A. Optical methods
2017, 8, 2275–2289. for monitoring neuron activity. Annu. Rev. Neurosci. 1978,
[107] Obien, M. E. J.; Deligkaris, K.; Bullmann, T.; Bakkum, 1, 171–182.
D. J.; Frey, U. Revealing neuronal function through [122] Grandy, T. H.; Greenfield, S. A.; Devonshire, I. M. An
microelectrode array recordings. Front. Neurosci. 2015, 8, evaluation of in vivo voltage-sensitive dyes: Pharmacological
423. side effects and signal-to-noise ratios after effective removal
[108] Seymour, J. P.; Wu, F.; Wise, K. D.; Yoon, E. State-of- of brain-pulsation artifacts. J. Neurophysiol. 2012, 108,
the-art MEMS and microsystem tools for brain research. 2931–2945.
Microsyst. Nanoeng. 2017, 3, 16066. [123] Nag, O. K.; Stewart, M. H.; Deschamps, J. R.; Susumu, K.;

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5101

Oh, E.; Tsytsarev, V.; Tang, Q. G.; Efros, A. L.; Vaxenburg, [135] Heim, N.; Garaschuk, O.; Friedrich, M. W.; Mank, M.;
R.; Black, B. J. et al. Quantum dot-peptide-fullerene Milos, R. I.; Kovalchuk, Y.; Konnerth, A.; Griesbeck, O.
bioconjugates for visualization of in vitro and in vivo cellular Improved calcium imaging in transgenic mice expressing
membrane potential. ACS Nano 2017, 11, 5598–5613. a troponin C-based biosensor. Nat. Methods 2007, 4,
[124] Nakai, J.; Ohkura, M.; Imoto, K. A high signal-to-noise 127–129.
Ca2+ probe composed of a single green fluorescent protein. [136] Palmer, A. E.; Giacomello, M.; Kortemme, T.; Hires, S.
Nat. Biotechnol. 2001, 19, 137–141. A.; Lev-Ram, V.; Baker, D.; Tsien, R. Y. Ca2+ indicators
[125] Nagai, T.; Yamada, S.; Tominaga, T.; Ichikawa, M.; based on computationally redesigned calmodulin-peptide
Miyawaki, A. Expanded dynamic range of fluorescent pairs. Chem. Biol. 2006, 13, 521–530.
indicators for Ca2+ by circularly permuted yellow fluorescent [137] Horikawa, K.; Yamada, Y.; Matsuda, T.; Kobayashi, K.;
proteins. Proc. Natl. Acad. Sci. USA 2004, 101, 10554– Hashimoto, M.; Matsu-ura, T.; Miyawaki, A.; Michikawa, T.;
10559. Mikoshiba, K.; Nagai, T. Spontaneous network activity
[126] Inoue, M.; Takeuchi, A.; Horigane, S.; Ohkura, M.; Gengyo- visualized by ultrasensitive Ca2+ indicators, yellow
Ando, K.; Fujii, H.; Kamijo, S.; Takemoto-Kimura, S.; Cameleon-Nano. Nat. Methods 2010, 7, 729–732.
Kano, M.; Nakai, J. et al. Rational design of a high-affinity, [138] Dreosti, E.; Odermatt, B.; Dorostkar, M. M.; Lagnado, L.
fast, red calcium indicator R-CaMP2. Nat. Methods 2015, A genetically encoded reporter of synaptic activity in vivo.
12, 64–70. Nat. Methods 2009, 6, 883–889.
[127] Akemann, W.; Mutoh, H.; Perron, A.; Rossier, J.; Knöpfel, [139] Zhao, Y. X.; Araki, S.; Wu, J. H.; Teramoto, T.; Chang, Y.
T. Imaging brain electric signals with genetically targeted F.; Nakano, M.; Abdelfattah, A. S.; Fujiwara, M.; Ishihara,
voltage-sensitive fluorescent proteins. Nat. Methods 2010, T.; Nagai, T. et al. An expanded palette of genetically
7, 643–649. encoded Ca2+ indicators. Science 2011, 333, 1888–1891.
[128] Miyawaki, A.; Llopis, J.; Heim, R.; McCaffery, J. M.; [140] Mank, M.; Santos, A. F.; Direnberger, S.; Mrsic-Flogel,
Adams, J. A.; Ikura, M.; Tsien, R. Y. Fluorescent T. D.; Hofer, S. B.; Stein, V.; Hendel, T.; Reiff, D. F.;
indicators for Ca2+ based on green fluorescent proteins and Levelt, C.; Borst, A. et al. A genetically encoded calcium
calmodulin. Nature 1997, 388, 882–887. indicator for chronic in vivo two-photon imaging. Nat.
[129] Burgoyne, R. D. Neuronal calcium sensor proteins: Methods 2008, 5, 805–811.
Generating diversity in neuronal Ca2+ signalling. Nat. Rev. [141] Xu, Y. X.; Zou, P.; Cohen, A. E. Voltage imaging with
Neurosci. 2007, 8, 182–193. genetically encoded indicators. Curr. Opin. Chem. Biol.
[130] Heim, N.; Griesbeck, O. Genetically encoded indicators of 2017, 39, 1–10.
cellular calcium dynamics based on troponin C and green [142] Jin, L.; Han, Z.; Platisa, J.; Wooltorton, J. R.; Cohen, L. B.;
fluorescent protein. J. Biol. Chem. 2004, 279, 14280–14286. Pieribone, V. A. Single action potentials and subthreshold
[131] Tian, L.; Hires, S. A.; Mao, T. Y.; Huber, D.; Chiappe, M. E.; electrical events imaged in neurons with a fluorescent
Chalasani, S. H.; Petreanu, L.; Akerboom, J.; McKinney, protein voltage probe. Neuron 2012, 75, 779–785.
S. A.; Schreiter, E. R. et al. Imaging neural activity in [143] Cao, G.; Platisa, J.; Pieribone, V. A.; Raccuglia, D.; Kunst,
worms, flies and mice with improved GCaMP calcium M.; Nitabach, M. N. Genetically targeted optical
indicators. Nat. Methods 2009, 6, 875–881. electrophysiology in intact neural circuits. Cell 2013, 154,
[132] Ahrens, M. B.; Orger, M. B.; Robson, D. N.; Li, J. M.; 904–913.
Keller, P. J. Whole-brain functional imaging at cellular [144] St-Pierre, F.; Marshall, J. D.; Yang, Y.; Gong, Y. Y.;
resolution using light-sheet microscopy. Nat. Methods Schnitzer, M. J.; Lin, M. Z. High-fidelity optical reporting
2013, 10, 413–420. of neuronal electrical activity with an ultrafast fluorescent
[133] Chen, T. W.; Wardill, T. J.; Sun, Y.; Pulver, S. R.; voltage sensor. Nat. Neurosci. 2014, 17, 884–889.
Renninger, S. L.; Baohan, A.; Schreiter, E. R.; Kerr, R. A.; [145] Lam, A. J.; St-Pierre, F.; Gong, Y. Y.; Marshall, J. D.;
Orger, M. B.; Jayaraman, V. et al. Ultrasensitive fluorescent Cranfill, P. J.; Baird, M. A.; McKeown, M. R.; Wiedenmann,
proteins for imaging neuronal activity. Nature 2013, 499, J.; Davidson, M. W.; Schnitzer, M. J. et al. Improving FRET
295–300. dynamic range with bright green and red fluorescent
[134] Barretto, R. P. J.; Gillis-Smith, S.; Chandrashekar, J.; proteins. Nat. Methods 2012, 9, 1005–1012.
Yarmolinsky, D. A.; Schnitzer, M. J.; Ryba, N. J. P.; Zuker, [146] Siegel, M. S.; Isacoff, E. Y. A genetically encoded optical
C. S. The neural representation of taste quality at the probe of membrane voltage. Neuron 1997, 19, 735–741.
periphery. Nature 2015, 517, 373–376. [147] Ataka, K.; Pieribone, V. A. A genetically targetable

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5102 Nano Res. 2018, 11(10): 5065–5106

fluorescent probe of channel gating with rapid kinetics. flexible and implantable microelectrode arrays using high-
Biophys. J. 2002, 82, 509–516. temperature grown vertical carbon nanotubes and a
[148] Sakai, R.; Repunte-Canonigo, V.; Raj, C. D.; Knöpfel, T. biocompatible polymer substrate. Nanotechnology 2015,
Design and characterization of a DNA-encoded, voltage- 26, 125301.
sensitive fluorescent protein. Eur. J. Neurosci. 2001, 13, [160] Panescu, D. Emerging technologies [wireless communication
2314–2318. systems for implantable medical devices]. IEEE Eng. Med.
[149] Guerrero, G.; Siegel, M. S.; Roska, B.; Loots, E.; Isacoff, Biol. Mag. 2008, 27, 96–101.
E. Y. Tuning FlaSh: Redesign of the dynamics, voltage [161] Beric, A.; Kelly, P. J.; Rezai, A.; Sterio, D.; Mogilner, A.;
range, and color of the genetically encoded optical sensor Zonenshayn, M.; Kopell, B. Complications of deep brain
of membrane potential. Biophys. J. 2002, 83, 3607–3618. stimulation surgery. Stereotact. Funct. Neurosurg. 2001,
[150] Baker, B. J.; Jin, L.; Han, Z.; Cohen, L. B.; Popovic, M.; 77, 73–78.
Platisa, J.; Pieribone, V. Genetically encoded fluorescent [162] Eom, K.; Hwang, S.; Yun, S.; Byun, K. M.; Jun, S. B.;
voltage sensors using the voltage-sensing domain of Kim, S. J. Photothermal activation of astrocyte cells using
Nematostella and Danio phosphatases exhibit fast kinetics. localized surface plasmon resonance of gold nanorods. J.
J. Neurosci. Methods 2012, 208, 190–196. Biophotonics 2017, 10, 486–493.
[151] Lundby, A.; Mutoh, H.; Dimitrov, D.; Akemann, W.; [163] Chen, R.; Romero, G.; Christiansen, M. G.; Mohr, A.;
Knöpfel, T. Engineering of a genetically encodable Anikeeva, P. Wireless magnetothermal deep brain stimulation.
fluorescent voltage sensor exploiting fast Ci-VSP voltage- Science 2015, 347, 1477–1480.
sensing movements. PLoS One 2008, 3, e2514. [164] Barolet, D. Light-emitting diodes (LEDs) in dermatology.
[152] Perron, A.; Mutoh, H.; Launey, T.; Knöpfel, T. Red-shifted Semin. Cutan. Med. Surg. 2008, 27, 227–238.
voltage-sensitive fluorescent proteins. Chem. Biol. 2009, [165] Legon, W.; Sato, T. F.; Opitz, A.; Mueller, J.; Barbour, A.;
16, 1268–1277. Williams, A.; Tyler, W. J. Transcranial focused ultrasound
[153] Akemann, W.; Mutoh, H.; Perron, A.; Park, Y. K.; Iwamoto, modulates the activity of primary somatosensory cortex in
Y.; Knöpfel, T. Imaging neural circuit dynamics with a humans. Nat. Neurosci. 2014, 17, 322–329.
voltage-sensitive fluorescent protein. J. Neurophysiol. [166] Deng, Z.-D.; Lisanby, S. H.; Peterchev, A. V. Electric
2012, 108, 2323–2337. field depth–focality tradeoff in transcranial magnetic
[154] Kralj, J. M.; Hochbaum, D. R.; Douglass, A. D.; Cohen, A. stimulation: Simulation comparison of 50 coil designs.
E. Electrical spiking in Escherichia coli probed with a Brain Stimul. 2013, 6, 1–13.
fluorescent voltage-indicating protein. Science 2011, 333, [167] Paviolo, C.; Haycock, J. W.; Cadusch, P. J.; McArthur,
345–348. S. L.; Stoddart, P. R. Laser exposure of gold nanorods can
[155] Kralj, J. M.; Douglass, A. D.; Hochbaum, D. R.; Maclaurin, induce intracellular calcium transients. J. Biophotonics
D.; Cohen, A. E. Optical recording of action potentials in 2014, 7, 761–765.
mammalian neurons using a microbial rhodopsin. Nat. [168] Choi, Y. K.; Lee, D. H.; Seo, Y. K.; Jung, H.; Park, J. K.;
Methods 2011, 9, 90–95. Cho, H. Stimulation of neural differentiation in human bone
[156] Hochbaum, D. R.; Zhao, Y. X.; Farhi, S. L.; Klapoetke, N.; marrow mesenchymal stem cells by extremely low-frequency
Werley, C. A.; Kapoor, V.; Zou, P.; Kralj, J. M.; electromagnetic fields incorporated with MNPs. Appl.
Maclaurin, D.; Smedemark-Margulies, N. et al. All-optical Biochem. Biotechnol. 2014, 174, 1233–1245.
electrophysiology in mammalian neurons using engineered [169] Nakatsuji, H.; Numata, T.; Morone, N.; Kaneko, S.; Mori, Y.;
microbial rhodopsins. Nat. Methods 2014, 11, 825–833. Imahori, H.; Murakami, T. Thermosensitive ion channel
[157] Wang, K.; Fishman, H. A.; Dai, H. J.; Harris, J. S. Neural activation in single neuronal cells by using surface-
stimulation with a carbon nanotube microelectrode array. engineered plasmonic nanoparticles. Angew. Chem., Int.
Nano Lett. 2006, 6, 2043–2048. Ed. 2015, 54, 11725–11729.
[158] Tsang, W. M.; Stone, A. L.; Otten, D.; Aldworth, Z. N.; [170] Bareket, L.; Waiskopf, N.; Rand, D.; Lubin, G.; David-Pur,
Daniel, T. L.; Hildebrand, J. G.; Levine, R. B.; Voldman, M.; Ben-Dov, J.; Roy, S.; Eleftheriou, C.; Sernagor, E.;
J. Insect-machine interface: A carbon nanotube-enhanced Cheshnovsky, O. et al. Semiconductor nanorod-carbon
flexible neural probe. J. Neurosci. Methods 2012, 204, nanotube biomimetic films for wire-free photostimulation
355–365. of blind retinas. Nano Lett. 2014, 14, 6685–6692.
[159] Yi, W. W.; Chen, C. Y.; Feng, Z. Y.; Xu, Y.; Zhou, C. P.; [171] Guduru, R.; Liang, P.; Hong, J.; Rodzinski, A.; Hadjikhani,
Masurkar, N.; Cavanaugh, J.; Ming-Cheng Cheng, M. A A.; Horstmyer, J.; Levister, E.; Khizroev, S. Magnetoelectric

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5103

“spin” on stimulating the brain. Nanomedicine 2015, 10, Nano Lett. 2007, 7, 513–519.
2051–2061. [186] Molokanova, E.; Bartel, J. A.; Zhao, W. W.; Naasani, I.;
[172] Marino, A.; Arai, S.; Hou, Y. Y.; Sinibaldi, E.; Pellegrino, Ignatius, M. J.; Treadway, J. A.; Savtchenko, A. Quantum
M.; Chang, Y. T.; Mazzolai, B.; Mattoli, V.; Suzuki, M.; dots move beyond fluorescence imaging. Biophot. Int.
Ciofani, G. Piezoelectric nanoparticle-assisted wireless 2008, 15, 26–31.
neuronal stimulation. ACS Nano 2015, 9, 7678–7689. [187] Yue, K.; Guduru, R.; Hong, J.; Liang, P.; Nair, M.; Khizroev,
[173] Carvalho-de-Souza, J. L.; Treger, J. S.; Dang, B. B.; Kent, S. Magneto-electric nano-particles for non-invasive brain
S. B. H.; Pepperberg, D. R.; Bezanilla, F. Photosensitivity stimulation. PLoS One 2012, 7, e44040.
of neurons enabled by cell-targeted gold nanoparticles. [188] Tyler, W. J.; Tufail, Y.; Finsterwald, M.; Tauchmann, M. L.;
Neuron 2015, 86, 207–217. Olson, E. J.; Majestic, C. Remote excitation of neuronal
[174] Shah, S.; Liu, J. J.; Pasquale, N.; Lai, J. P.; McGowan, H.; circuits using low-intensity, low-frequency ultrasound.
Pang, Z. P.; Lee, K. B. Hybrid upconversion nanomaterials PLoS One 2008, 3, e3511.
for optogenetic neuronal control. Nanoscale 2015, 7, [189] Wang, Y. C.; Wu, Y.; Quadri, F.; Prox, J. D.; Guo, L.
16571–16577. Cytotoxicity of ZnO nanowire arrays on excitable cells.
[175] Tay, A.; Kunze, A.; Murray, C.; Di Carlo, D. Induction of Nanomaterial 2017, 7, 80.
calcium influx in cortical neural networks by nanomagnetic [190] Ciofani, G.; Danti, S.; D’Alessandro, D.; Ricotti, L.;
forces. ACS Nano 2016, 10, 2331–2341. Moscato, S.; Bertoni, G.; Falqui, A.; Berrettini, S.; Petrini,
[176] Catterall, W. A. Structure and function of voltage-gated M.; Mattoli, V. et al. Enhancement of neurite outgrowth in
ion channels. Annu. Rev. Biochem. 1995, 64, 493–531. neuronal-like cells following boron nitride nanotube-
[177] Bean, B. P. The action potential in mammalian central mediated stimulation. ACS Nano 2010, 4, 6267–6277.
neurons. Nat. Rev. Neurosci. 2007, 8, 451–465. [191] Ricotti, L.; Fujie, T.; Vazão, H.; Ciofani, G.; Marotta, R.;
[178] Bareket-Keren, L.; Hanein, Y. Novel interfaces for light Brescia, R.; Filippeschi, C.; Corradini, I.; Matteoli, M.;
directed neuronal stimulation: Advances and challenges. Mattoli, V. et al. Boron nitride nanotube-mediated stimulation
Int. J. Nanomedicine 2014, 9, 65–83. of cell co-culture on micro-engineered hydrogels. PLoS
[179] Smith, A. M.; Mancini, M. C.; Nie, S. M. Second window One 2013, 8, e71707.
for in vivo imaging. Nat. Nanotechnol. 2009, 4, 710–711. [192] Genchi, G. G.; Ceseracciu, L.; Marino, A.; Labardi, M.;
[180] Del Bonis-O’Donnell, J. T.; Page, R. H.; Beyene, A. G.; Marras, S.; Pignatelli, F.; Bruschini, L.; Mattoli, V.;
Tindall, E. G.; McFarlane, I. R.; Landry, M. P. Dual Ciofani, G. P(VDF-TrFE)/BaTiO3 nanoparticle composite
near-infrared two-photon microscopy for deep-tissue films mediate piezoelectric stimulation and promote
dopamine nanosensor imaging. Adv. Funct. Mater. 2017, differentiation of SH-SY5Y neuroblastoma cells. Adv.
27, 1702112. Healthc. Mater. 2016, 5, 1808–1820.
[181] Lugo, K.; Miao, X. Y.; Rieke, F.; Lin, L. Y. Remote [193] Rojas, C.; Tedesco, M.; Massobrio, P.; Marino, A.;
switching of cellular activity and cell signaling using light Ciofani, G.; Martinoia, S.; Raiteri, R. Acoustic stimulation
in conjunction with quantum dots. Biomed. Opt. Express can induce a selective neural network response mediated
2012, 3, 447–454. by piezoelectric nanoparticles. J. Neural Eng. 2018, 15,
[182] Gomez, N.; Winter, J. O.; Shieh, F.; Saunders, A. E.; 036016.
Korgel, B. A.; Schmidt, C. E. Challenges in quantum [194] Marino, A.; Barsotti, J.; de Vito, G.; Filippeschi, C.; Mazzolai,
dot-neuron active interfacing. Talanta 2005, 67, 462–471. B.; Piazza, V.; Labardi, M.; Mattoli, V.; Ciofani, G. Two-
[183] Winter, J. O.; Liu, T. Y.; Korgel, B. A.; Schmidt, C. E. photon lithography of 3D nanocomposite piezoelectric
Recognition molecule directed interfacing between scaffolds for cell stimulation. ACS Appl. Mater. Interfaces
semiconductor quantum dots and nerve cells. Adv. Mater. 2015, 7, 25574–25579.
2001, 13, 1673–1677. [195] Hoop, M.; Chen, X. Z.; Ferrari, A.; Mushtaq, F.;
[184] Winter, J. O.; Gomez, N.; Korgel, B. A.; Schmidt, C. E. Ghazaryan, G.; Tervoort, T.; Poulikakos, D.; Nelson, B.;
Quantum dots for electrical stimulation of neural cells. Pané, S. Ultrasound-mediated piezoelectric differentiation
Proceedings of SPIE 2005, 5705, 235–246. of neuron-like PC12 cells on PVDF membranes. Sci. Rep.
[185] Pappas, T. C.; Wickramanyake, W. M. S.; Jan, E.; 2017, 7, 4028.
Motamedi, M.; Brodwick, M.; Kotov, N. A. Nanoscale [196] Lee, Y.-S.; Collins, G.; Arinzeh, T. L. Neurite extension
engineering of a cellular interface with semiconductor of primary neurons on electrospun piezoelectric scaffolds.
nanoparticle films for photoelectric stimulation of neurons. Acta Biomater. 2011, 7, 3877–3886.

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5104 Nano Res. 2018, 11(10): 5065–5106

[197] Eustis, S.; El-Sayed, M. A. Why gold nanoparticles are Science 2012, 336, 604–608.
more precious than pretty gold: Noble metal surface [210] Stanley, S. A.; Sauer, J.; Kane, R. S.; Dordick, J. S.;
plasmon resonance and its enhancement of the radiative Friedman, J. M. Remote regulation of glucose homeostasis
and nonradiative properties of nanocrystals of different in mice using genetically encoded nanoparticles. Nat.
shapes. Chem. Soc. Rev. 2006, 35, 209–217. Med. 2015, 21, 92–98.
[198] Shapiro, M. G.; Homma, K.; Villarreal, S.; Richter, C. P.; [211] Boyden, E. S.; Zhang, F.; Bamberg, E.; Nagel, G.;
Bezanilla, F. Infrared light excites cells by changing their Deisseroth, K. Millisecond-timescale, genetically targeted
electrical capacitance. Nat. Commun. 2012, 3, 736. optical control of neural activity. Nat. Neurosci. 2005, 8,
[199] Benham, C. D.; Gunthorpe, M. J.; Davis, J. B. TRPV 1263–1268.
channels as temperature sensors. Cell Calcium 2003, 33, [212] Lin, J. Y.; Knutsen, P. M.; Muller, A.; Kleinfeld, D.; Tsien,
479–487. R. Y. ReaChR: A red-shifted variant of channelrhodopsin
[200] Paviolo, C.; Haycock, J. W.; Yong, J.; Yu, A.; Stoddart, P. enables deep transcranial optogenetic excitation. Nat.
R.; McArthur, S. L. Laser exposure of gold nanorods can Neurosci. 2013, 16, 1499–1508.
increase neuronal cell outgrowth. Biotechnol. Bioeng. [213] Klapoetke, N. C.; Murata, Y.; Kim, S. S.; Pulver, S. R.;
2013, 110, 2277–2291. Birdsey-Benson, A.; Cho, Y. K.; Morimoto, T. K.; Chuong,
[201] Yong, J.; Needham, K.; Brown, W. G. A.; Nayagam, A. S.; Carpenter, E. J.; Tian, Z. J. et al. Independent optical
B. A.; McArthur, S. L.; Yu, A. M.; Stoddart, P. R. excitation of distinct neural populations. Nat. Methods
Gold-nanorod-assisted near-infrared stimulation of primary 2014, 11, 338–346.
auditory neurons. Adv. Healthc. Mater. 2014, 3, 1862–1868. [214] Chuong, A. S.; Miri, M. L.; Busskamp, V.; Matthews, G.
[202] Eom, K.; Kim, J.; Choi, J. M.; Kang, T.; Chang, J. W.; A.; Acker, L. C.; Sørensen, A. T.; Young, A.; Klapoetke,
Byun, K. M.; Jun, S. B.; Kim, S. J. Enhanced infrared N. C.; Henninger, M. A.; Kodandaramaiah, S. B. et al.
neural stimulation using localized surface plasmon resonance Noninvasive optical inhibition with a red-shifted microbial
of gold nanorods. Small 2014, 10, 3853–3857. rhodopsin. Nat. Neurosci. 2014, 17, 1123–1129.
[203] Yoo, S.; Hong, S.; Choi, Y.; Park, J. H.; Nam, Y. [215] Pansare, V. J.; Hejazi, S.; Faenza, W. J.; Prud’homme, R.
Photothermal inhibition of neural activity with near- K. Review of long-wavelength optical and NIR imaging
infrared-sensitive nanotransducers. ACS Nano 2014, 8, materials: Contrast agents, fluorophores, and multifunctional
8040–8049. nano carriers. Chem. Mater. 2012, 24, 812–827.
[204] Lavoie-Cardinal, F.; Salesse, C.; Bergeron, É.; Meunier, [216] Hososhima, S.; Yuasa, H.; Ishizuka, T.; Hoque, M. R.;
M.; De Koninck, P. Gold nanoparticle-assisted all optical Yamashita, T.; Yamanaka, A.; Sugano, E.; Tomita, H.;
localized stimulation and monitoring of Ca2+ signaling in Yawo, H. Near-infrared (NIR) up-conversion optogenetics.
neurons. Sci. Rep. 2016, 6, 20619. Sci. Rep. 2015, 5, 16533.
[205] Eom, K.; Im, C.; Hwang, S.; Eom, S.; Kim, T. S.; Jeong, [217] Bansal, A.; Liu, H. C.; Jayakumar, M. K. G.; Andersson-
H. S.; Kim, K. H.; Byun, K. M.; Jun, S. B.; Kim, S. J. Engels, S.; Zhang, Y. Quasi-continuous wave near-infrared
Synergistic combination of near-infrared irradiation and excitation of upconversion nanoparticles for optogenetic
targeted gold nanoheaters for enhanced photothermal neural manipulation of C. elegans. Small 2016, 12, 1732–1743.
stimulation. Biomed. Opt. Express 2016, 7, 1614–1625. [218] He, L.; Zhang, Y. W; Ma, G. L.; Tan, P.; Li, Z. J.; Zang,
[206] Yoo, S.; Kim, R.; Park, J. H.; Nam, Y. Electro-optical S. B.; Wu, X.; Jing, J.; Fang, S. H.; Zhou, L. J. et al.
neural platform integrated with nanoplasmonic inhibition Near-infrared photoactivatable control of Ca2+ signaling
interface. ACS Nano 2016, 10, 4274–4281. and optogenetic immunomodulation. eLife 2015, 4, e10024.
[207] Bazard, P.; Frisina, R. D.; Walton, J. P.; Bhethanabotla, V. [219] Wu, X.; Zhang, Y. W.; Takle, K.; Bilsel, O.; Li, Z. J.; Lee,
R. Nanoparticle-based plasmonic transduction for modulation H.; Zhang, Z. J.; Li, D. S.; Fan, W.; Duan, C. Y. et al.
of electrically excitable cells. Sci. Rep. 2017, 7, 7803. Dye-sensitized core/active shell upconversion nanoparticles
[208] Tay, A.; Di Carlo, D. Magnetic nanoparticle-based mechanical for optogenetics and bioimaging applications. ACS Nano
stimulation for restoration of mechano-sensitive ion 2016, 10, 1060–1066.
channel equilibrium in neural networks. Nano Lett. 2017, [220] Huang, K.; Dou, Q. Q.; Loh, X. J. Nanomaterial mediated
17, 886–892. optogenetics: Opportunities and challenges. RSC Adv. 2016,
[209] Stanley, S. A.; Gagner, J. E.; Damanpour, S.; Yoshida, M.; 6, 60896–60906.
Dordick, J. S.; Friedman, J. M. Radio-wave heating of iron [221] El Haj, A. J.; Hughes, S.; Dobson, J. Manipulation of ion
oxide nanoparticles can regulate plasma glucose in mice. channels using magnetic micro- and nanoparticle cytometry.

| www.editorialmanager.com/nare/default.asp
Nano Res. 2018, 11(10): 5065–5106 5105

Comp. Biochem. Phys. A 2003, 134, S110. [235] Tay, A. K.; Dhar, M.; Pushkarsky, I.; Di Carlo, D. Research
[222] Hughes, S.; El Haj, A. J.; Dobson, J. Magnetic micro- and highlights: Manipulating cells inside and out. Lab Chip
nanoparticle mediated activation of mechanosensitive ion 2015, 15, 2533–2537.
channels. Med. Eng. Phys. 2005, 27, 754–762. [236] Miesenböck, G. Optogenetic control of cells and circuits.
[223] Kunze, A.; Tseng, P.; Godzich, C.; Murray, C.; Caputo, A.; Annu. Rev. Cell. Dev. Biol. 2011, 27, 731–758.
Schweizer, F. E.; Di Carlo, D. Engineering cortical neuron [237] Deisseroth, K. Optogenetics: 10 years of microbial opsins
polarity with nanomagnets on a chip. ACS Nano 2015, 9, in neuroscience. Nat. Neurosci. 2015, 18, 1213–1225.
3664–3676. [238] Boyden, E. S. Optogenetics and the future of neuroscience.
[224] Etoc, F.; Vicario, C.; Lisse, D.; Siaugue, J. M.; Piehler, J.; Nat. Neurosci. 2015, 18, 1200–1201.
Coppey, M.; Dahan, M. Magnetogenetic control of protein [239] Boyden, E. S. A history of optogenetics: The development
gradients inside living cells with high spatial and temporal of tools for controlling brain circuits with light. F1000 Biol.
resolution. Nano Lett. 2015, 15, 3487–3494. Rep. 2011, 3, 11.
[225] Wang, N.; Butler, J. P.; Ingber, D. E. Mechanotransduction [240] Knöpfel, T.; Boyden, E. S. Optogenetics: Tools for
across the cell surface and through the cytoskeleton. Science controlling and monitoring neuronal activity preface. Prog.
Brain Res. 2012, 196, VII–VIII.
1993, 260, 1124–1127.
[241] Deisseroth, K. Optogenetics. Nat. Methods 2011, 8, 26–29.
[226] Wang, N.; Ingber, D. E. Control of cytoskeletal mechanics
[242] Gerits, A.; Vanduffel, W. Optogenetics in primates: A
by extracellular matrix, cell shape, and mechanical tension.
shining future? Trends Genet. 2013, 29, 403–411.
Biophys. J. 1994, 66, 1281–1289.
[243] Madisen, L.; Mao, T. Y.; Koch, H.; Zhuo, J. M.; Berenyi, A.;
[227] Wang, N.; Ingber, D. E. Probing transmembrane mechanical
Fujisawa, S.; Hsu, Y. W. A.; Garcia, A. J., 3rd; Gu, X.;
coupling and cytomechanics using magnetic twisting
Zanella, S. et al. A toolbox of Cre-dependent optogenetic
cytometry. Biochem. Cell Biol. 1995, 73, 327–335.
transgenic mice for light-induced activation and silencing.
[228] Glogauer, M.; Ferrier, J.; McCulloch, C. A. Magnetic
Nat. Neurosci. 2012, 15, 793–802.
fields applied to collagen-coated ferric oxide beads induce
[244] Zhang, F.; Vierock, J.; Yizhar, O.; Fenno, L. E.; Tsunoda,
stretch-activated Ca2+ flux in fibroblasts. Am. J. Physiol.
S.; Kianianmomeni, A.; Prigge, M.; Berndt, A.; Cushman, J.;
1995, 269, C1093–1104.
Polle, J. et al. The microbial opsin family of optogenetic
[229] Hughes, S.; McBain, S.; Dobson, J.; El Haj, A. J. Selective
tools. Cell 2011, 147, 1446–1457.
activation of mechanosensitive ion channels using magnetic
[245] Gautier, A.; Gauron, C.; Volovitch, M.; Bensimon, D.;
particles. J. R. Soc. Interface 2008, 5, 855–863. Jullien, L.; Vriz, S. How to control proteins with light in
[230] Matthews, B. D.; Thodeti, C. K.; Tytell, J. D.; Mammoto, living systems. Nat. Chem. Biol. 2014, 10, 533–541.
A.; Overby, D. R.; Ingber, D. E. Ultra-rapid activation [246] Land, B. B.; Brayton, C. E.; Furman, K. E.; Lapalombara, Z.;
of TRPV4 ion channels by mechanical forces applied to DiLeone, R. J. Optogenetic inhibition of neurons by internal
cell surface β1 integrins. Integr. Biol. 2010, 2, 435–442. light production. Front. Behav. Neurosci. 2014, 8, 108.
[231] Mannix, R. J.; Kumar, S.; Cassiola, F.; Montoya-Zavala, M.; [247] Nihongaki, Y.; Kawano, F.; Nakajima, T.; Sato, M.
Feinstein, E.; Prentiss, M.; Ingber, D. E. Nanomagnetic Photoactivatable CRISPR-Cas9 for optogenetic genome
actuation of receptor-mediated signal transduction. Nat. editing. Nat. Biotechnol. 2015, 33, 755–760.
Nanotechnol. 2008, 3, 36–40. [248] Nihongaki, Y.; Yamamoto, S.; Kawano, F.; Suzuki, H.;
[232] Cho, M. H.; Lee, E. J.; Son, M.; Lee, J. H.; Yoo, D.; Kim, Sato, M. CRISPR-Cas9-based photoactivatable transcription
J.; Park, S. W.; Shin, J. S.; Cheon, J. A magnetic switch system. Chem. Biol. 2015, 22, 169–174.
for the control of cell death signalling in in vitro and in [249] Gradinaru, V.; Zhang, F.; Ramakrishnan, C.; Mattis, J.;
vivo systems. Nat. Mater. 2012, 11, 1038–1043. Prakash, R.; Diester, I.; Goshen, I.; Thompson, K. R.;
[233] Bharde, A. A.; Palankar, R.; Fritsch, C.; Klaver, A.; Deisseroth, K. Molecular and cellular approaches for
Kanger, J. S.; Jovin, T. M.; Arndt-Jovin, D. J. Magnetic diversifying and extending optogenetics. Cell 2010, 141,
nanoparticles as mediators of ligand-free activation of 154–165.
EGFR signaling. PLoS One 2013, 8, e68879. [250] Zhang, F.; Wang, L. P.; Brauner, M.; Liewald, J. F.;
[234] Steketee, M. B.; Moysidis, S. N.; Jin, X. L.; Weinstein, J. E.; Kay, K.; Watzke, N.; Wood, P. G.; Bamberg, E.; Nagel, G.;
Pita-Thomas, W.; Raju, H. B.; Iqbal, S.; Goldberg, J. L. Gottschalk, A. et al. Multimodal fast optical interrogation
Nanoparticle-mediated signaling endosome localization of neural circuitry. Nature 2007, 446, 633–639.
regulates growth cone motility and neurite growth. Proc. [251] Gradinaru, V.; Thompson, K. R.; Deisseroth, K. eNpHR:
Natl. Acad. Sci. USA 2011, 108, 19042–19047. A natronomonas halorhodopsin enhanced for optogenetic

www.theNanoResearch.com∣www.Springer.com/journal/12274 | Nano Research


5106 Nano Res. 2018, 11(10): 5065–5106

applications. Brain Cell Biol. 2008, 36, 129–139. magnetic nanoparticles. Nat. Nanotechnol. 2008, 3, 139–143.
[252] Zhao, S. L.; Cunha, C.; Zhang, F.; Liu, Q.; Gloss, B.; [264] Monzel, C.; Vicario, C.; Piehler, J.; Coppey, M.; Dahan, M.
Deisseroth, K.; Augustine, G. J.; Feng, G. P. Improved Magnetic control of cellular processes using biofunctional
expression of halorhodopsin for light-induced silencing of nanoparticles. Chem. Sci. 2017, 8, 7330–7338.
neuronal activity. Brain Cell Biol. 2008, 36, 141–154. [265] Patel, A. J.; Honoré, E. Properties and modulation of
[253] Zhang, F.; Prigge, M.; Beyrière, F.; Tsunoda, S. P.; mammalian 2P domain K+ channels. Trends Neurosci.
Mattis, J.; Yizhar, O.; Hegemann, P.; Deisseroth, K. 2001, 24, 339–346.
Red-shifted optogenetic excitation: A tool for fast neural [266] Wheeler, M. A.; Smith, C. J.; Ottolini, M.; Barker, B. S.;
control derived from Volvox carteri. Nat. Neurosci. 2008, Purohit, A. M.; Grippo, R. M.; Gaykema, R. P.; Spano, A. J.;
11, 631–633. Beenhakker, M. P.; Kucenas, S. et al. Genetically targeted
[254] Yizhar, O.; Fenno, L. E.; Prigge, M.; Schneider, F.; magnetic control of the nervous system. Nat. Neurosci.
Davidson, T. J.; O’Shea, D. J.; Sohal, V. S.; Goshen, I.; 2016, 19, 756–761.
Finkelstein, J.; Paz, J. T. et al. Neocortical excitation/ [267] McKemy, D. D.; Neuhausser, W. M.; Julius, D. Identification
inhibition balance in information processing and social of a cold receptor reveals a general role for TRP channels
dysfunction. Nature 2011, 477, 171–178. in thermosensation. Nature 2002, 416, 52–58.
[255] Chow, B. Y.; Han, X.; Dobry, A. S.; Qian, X. F.; Chuong, [268] Stanley, S. A.; Kelly, L.; Latcha, K. N.; Schmidt, S. F.; Yu,
A. S.; Li, M. J.; Henninger, M. A.; Belfort, G. M.; Lin, Y. X. F.; Nectow, A. R.; Sauer, J.; Dyke, J. P.; Dordick, J. S.;
X.; Monahan, P. E. et al. High-performance genetically Friedman, J. M. Bidirectional electromagnetic control of the
targetable optical neural silencing by light-driven proton hypothalamus regulates feeding and metabolism. Nature
pumps. Nature 2010, 463, 98–102. 2016, 531, 647–650.
[256] Mattis, J.; Tye, K. M.; Ferenczi, E. A.; Ramakrishnan, C.; [269] Qin, S. Y.; Yin, H.; Yang, C. L.; Dou, Y. F.; Liu, Z. M.;
O'Shea, D. J.; Prakash, R.; Gunaydin, L. A.; Hyun, M.; Zhang, P.; Yu, H.; Huang, Y. L.; Feng, J.; Hao, J. F. et al.
Fenno, L. E.; Gradinaru, V. et al. Principles for applying A magnetic protein biocompass. Nat. Mater. 2016, 15,
optogenetic tools derived from direct comparative analysis 217–226.
of microbial opsins. Nat. Methods 2011, 9, 159–172. [270] Long, X. Y.; Ye, J.; Zhao, D.; Zhang, S. J. Magnetogenetics:
[257] Wietek, J.; Wiegert, J. S.; Adeishvili, N.; Schneider, F.; Remote non-invasive magnetic activation of neuronal activity
Watanabe, H.; Tsunoda, S. P.; Vogt, A.; Elstner, M.; Oertner, with a magnetoreceptor. Sci. Bull. 2015, 60, 2107–2119.
T. G.; Hegemann, P. Conversion of channelrhodopsin into a [271] Ibsen, S.; Tong, A.; Schutt, C.; Esener, S.; Chalasani, S. H.
light-gated chloride channel. Science 2014, 344, 409–412. Sonogenetics is a non-invasive approach to activating
[258] Govorunova, E. G.; Sineshchekov, O. A.; Janz, R.; Liu, X. Q.; neurons in Caenorhabditis elegans. Nat. Commun. 2015,
Spudich, J. L. Natural light-gated anion channels: A family 6, 8264.
of microbial rhodopsins for advanced optogenetics. Science [272] Kubanek, J.; Shi, J. Y.; Marsh, J.; Chen, D.; Deng, C. R.;
2015, 349, 647–650. Cui, J. M. Ultrasound modulates ion channel currents. Sci.
[250] Lin, J. Y.; Lin, M. Z.; Steinbach, P.; Tsien, R. Y. Rep. 2016, 6, 24170.
Characterization of engineered channelrhodopsin variants [273] Spira, M. E.; Hai, A. Multi-electrode array technologies
with improved properties and kinetics. Biophys. J. 2009, for neuroscience and cardiology. Nat. Nanotechnol. 2013,
96, 1803–1814. 8, 83–94.
[260] Berndt, A.; Schoenenberger, P.; Mattis, J.; Tye, K. M.; [274] Berger, T. W.; Baudry, M.; Brinton, R. D.; Liaw, J.-S.;
Deisseroth, K.; Hegemann, P.; Oertner, T. G. High- Marmarelis, V. Z.; Park, A. Y.; Sheu, B. J.; Tanguay, A. R.
efficiency channelrhodopsins for fast neuronal stimulation Brain-implantable biomimetic electronics as the next era
at low light levels. Proc. Natl. Acad. Sci. USA 2011, 108, in neural prosthetics. Proc. IEEE 2001, 89, 993–1012.
7595–7600. [275] Berger, T. W.; Hampson, R. E.; Song, D.; Goonawardena, A.;
[261] Matsuzaki, M.; Ellis-Davies, G. C.; Nemoto, T.; Miyashita, Marmarelis, V. Z.; Deadwyler, S. A. A cortical neural
Y.; Iino, M.; Kasai, H. Dendritic spine geometry is critical prosthesis for restoring and enhancing memory. J. Neural
for AMPA receptor expression in hippocampal CA1 Eng. 2011, 8, 046017.
pyramidal neurons. Nat. Neurosci. 2001, 4, 1086–1092. [276] Ezzyat, Y.; Wanda, P. A.; Levy, D. F.; Kadel, A.; Aka, A.;
[262] Gorostiza, P.; Isacoff, E. Y. Optical switches for remote Pedisich, I.; Sperling, M. R.; Sharan, A. D.; Lega, B. C.;
and noninvasive control of cell signaling. Science 2008, Burks, A. et al. Closed-loop stimulation of temporal cortex
322, 395–399. rescues functional networks and improves memory. Nat.
[263] Dobson, J. Remote control of cellular behaviour with Commun. 2018, 9, 365.

| www.editorialmanager.com/nare/default.asp

You might also like