Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

REVIEWS

Neural recording and modulation


technologies
Ritchie Chen, Andres Canales and Polina Anikeeva
Abstract | In the mammalian nervous system, billions of neurons connected by quadrillions of
synapses exchange electrical, chemical and mechanical signals. Disruptions to this network
manifest as neurological or psychiatric conditions. Despite decades of neuroscience research,
our ability to treat or even to understand these conditions is limited by the capability of tools to
probe the signalling complexity of the nervous system. Although orders of magnitude smaller and
computationally faster than neurons, conventional substrate-bound electronics do not
recapitulate the chemical and mechanical properties of neural tissue. This mismatch results in a
foreign-body response and the encapsulation of devices by glial scars, suggesting that the design
of an interface between the nervous system and a synthetic sensor requires additional materials
innovation. Advances in genetic tools for manipulating neural activity have fuelled the demand
for devices that are capable of simultaneously recording and controlling individual neurons at
unprecedented scales. Recently, flexible organic electronics and bio- and nanomaterials have
been developed for multifunctional and minimally invasive probes for long-term interaction with
the nervous system. In this Review, we discuss the design lessons from the quarter-century-old
field of neural engineering, highlight recent materials-driven progress in neural probes and look
at emergent directions inspired by the principles of neural transduction.

Understanding the information transfer and processing which include astrocytes, oligodendrocytes and micro-
within the mammalian nervous system is one of the most glia7; the roles of these cells in brain function remains
urgent challenges faced by the biomedical community. an active area of research8,9. Similarly, the ~45‑cm-long,
Neurological, neurodegenerative, psychiatric and neu- 1.5‑cm-thick spinal cord contains a range of motor and
romuscular conditions, for example, Parkinson disease, interneurons, Schwann cells and neuronal processes
Alzheimer disease, major depressive disorder and mul- that connect the brain to the >150,000‑km web of the
tiple sclerosis, respectively, affect an ever-increasing peripheral nervous system10 (FIG. 1g–i).
percentage of our ageing population1–3 (FIG. 1a–b), and Neural activity is marked by millisecond-long
traumatic injuries to the nervous system contribute 80–100‑mV spikes in cell-membrane voltage called
2.4 million patients each year to the disability burden action potentials6. Aided by voltage-sensitive ion chan-
in the United States alone4. Furthermore, inflammatory nels, action potentials propagate across neuronal mem-
conditions, such as hypertension and arthritis, have been branes and trigger the release of neuro­transmitters
linked to changes in peripheral nerve activity 5. Although from pre­synaptic terminals into the synaptic cleft.
the basic neuroscience research of the past century has Neurotransmitters can then bind to and activate recep-
greatly advanced our understanding of neuronal func- tors on the postsynaptic neuronal membrane, which
Department of Materials tion, the ability to record and manipulate the dynamics results in signal transduction. In addition to chemical
Science and Engineering, of the nervous system remains insufficient to treat these and electrical signals, neurons can respond to physical
and Research Laboratory of psychiatric and neurological disorders or to restore stimuli, such as pH11, temperature12,13, and pressure and
Electronics, Massachusetts
function following nerve injury. tension14, through various ion channels. Understanding
Institute of Technology,
Cambridge, Massachusetts Within the 1.3‑litre volume of the human brain, bil- neural function hence requires tools that can commu-
02193, USA. lions of neurons, divided into thousands of genetically nicate with neurons across the diverse range of their
Correspondence to P.A. and structurally defined subtypes, communicate with signalling modalities.
anikeeva@mit.edu each other through quadrillions of synapses6 (FIG. 1c–f). Driven by Moore’s law over the past 25 years, minia-
doi:10.1038/natrevmats.2016.93 Notably, neurons constitute only ~50% of the cells in turization of electronic devices down to the nanoscale has
Published online 4 Jan 2017 the brain. The other half are electrically inactive glia, delivered unprecedented computational power. Although

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 1


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a b
5 4.96 215
Patients with Parkinson disease (million)

210
4 210

Annual cost (billion USD)


205
3

15 14.5 14.4
13.6
2 1.97
10
1.22
1 0.82 0.87
0.61 0.70 5
0.35 0.49
0.33 0.33
0.16
0 0
China

EU

US

India

Brazil

Others

MS

SCIs

Parkinson
disease

Depression
c 16.7 cm d e f
Excitation Axon
Inhibition
Cortex Myelin sheath
9.3 cm

From
Thalamus SNc D1 D2 Dendrite
GPi
Blood vessel GPe
STN
Putamen
Synapse
Spinal cord

Neuron Microglial cell • Direct path way activates • 1011 neurons


movement • 1014 synapses
• Indirect pathway inhibits • 102 neurotransmitters
movement • Neuron size 1–20 μm
• Firing rates 0.1–200 Hz
• Elastic modulus ~kPa
g h i
45 cm

Central canal
Dura mater

1–1.5 cm Extensor Flexor


White matter Spinal nerve Inhibitory interneuron • 109 neurons
Renshaw cell • >150,000 km of nerve fibres
Grey matter
Motor neurons • 103–104 nerve fibres in cross-section
Sensory afferent • Myelinated and unmyelinated axons
• Modulus ~kPa–MPa
• Repeated strain ~12%

Figure 1 | Clinical impact of the diseases and injuries of the nervous system and physiological guidelines
Nature Reviewsfor| Materials
designing neural interfaces. a | The worldwide population of diagnosed Parkinson disease cases in 2005 (4.1 million) and
the corresponding projection to 2030 (8.7 million)1. b | Combined annual cost of multiple sclerosis (MS)2, spinal cord injuries
(SCIs)212, Parkinson disease213 and depression3 in the United States displayed in US$ billions. c | Dimensions of the brain and
spinal cord. d | Brain tissue comprising neurons, glia and blood vessels. e | Circuit diagram for direct and indirect motor
pathways in the brain214. f | Physiological facts pertaining to brain function. g | A cross-section of the spinal cord. h | Circuit
diagram of the motor pathway in the spinal cord215,216. i | Cross-sectional diagram of a peripheral nerve, and physiological
facts pertaining to the functions of the spinal cord and peripheral nerves10. D1 and D2 are dopamine receptors. GPe,
external globus pallidus; GPi, internal globus pallidus; SNc, substantia nigra pars compacta; STN, subthalamic nucleus.

2 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Genetic tetrodes19,20 were developed to investigate communi-


Rhodopsins ChR2 expression in neurons cation between large groups of neurons within distinct
neural circuits.
These innovations in probe design relied on tailor-
1960s 1970s 1990s 2000s 2010s ing the device geometry and electrical impedance to a
particular neuroscience application. For extracellular
recordings, the impedance is determined by the capac-
Green fluorescent protein Fluorescent ion indicators CRISPR–Cas9 gene editing
itive characteristics of the interface between the elec-
trode and the cerebrospinal fluid, which depends on
b Electrical the electrode material and the tip area. High-impedance
electrodes (1–5 MΩ) are typically used to record action
Synaptic transmission Brain–machine interfaces potentials with high signal-to‑noise ratios (SNR > 20)
from neurons within a 10‑μm radius21. Insulated steel,
Electrical signalling Circuits neuroscience Systems neuroscience
tungsten, gold and platinum microwires with dimensions
of ~10–100 μm and impedances of <1 MΩ are commonly
1950s 1970s 1980s 1990s 2000s used for extracellular recordings from moderate num-
bers of neurons22. Stereotrodes and tetrodes consist of
two or four polymer-insulated 12‑μm nickel–chromium
Patch clamp Arrays: Utah, Michigan Miniaturization microwires that are electrochemically coated with gold to
and tetrodes and refinement lower the impedance from ~1–3 MΩ to 100–500 kΩ and
to improve biochemical stability 19,20. The close proximity
c Materials of multiple microwires enables deconvolution of over-
Interactions of central Whole-brain, spinal lapping population signals to identify action-potential
Long-term and peripheral Complex cord, peripheral shapes corresponding to specific cells via principal com-
failure modes nervous systems behaviour nerve interfaces ponent analysis23. Neurotrophic electrodes integrate a
microwire within a pipette carrying neurotrophic fac-
tors that stimulate neuronal ingrowth for high-fidelity
2000s 2010s 2020s
recordings24.

Coatings on Flexible electronics Wireless platforms Nanotransducers Microfabricated neural probes. By the 1980s, advances in
hard probes semiconductor microfabrication had led to the develop­
ment of silicon-based multielectrode arrays known as
Figure 2 | Historical view of neural interface research. a | Progress in genetic tools for Utah arrays17 and Michigan probes18. Utah arrays, which
imaging and manipulating neural circuits paralleled with engineering trends in|neural
Nature Reviews Materials contain up to 128 sharp, metal-tipped electrodes with
probe design. b | The development of neural interfaces, driven by the needs of brain–
a pitch of 200–400 μm, are produced from thick silicon
machine interface research, has been confounded by the failure of neural probes in
chronic long-term experiments. c | Recently, there has been a materials-driven pursuit to wafers by a combination of micromachining and litho­
reduce the foreign-body response, push the resolution of neural interfaces to subcellular graphy. Implanted Utah arrays ‘float’ on the surface of
dimensions and integrate the capabilities required for the deployment of genetic tools. the brain and are connected to skull-mounted interface
boards by flexible cables. Owing to their relatively large
area and electrode count, these devices have become
neuroprosthetic research has undoubtedly benefited from crucial components in studies of cortical circuits in non-
these advances, additional design parameters need to be human primates and are the only neural probes approved
included for effective long-term operation and clinical by the US Food and Drug Administration for chronic
translation of neural interfaces. The complexity of neural use in human patients25. As Utah arrays only permit
tissue necessitates further materials innovation beyond recordings from the topmost 1–3 mm of the cortex and
microfabricated semiconductor circuits. Consideration because large footprints limit their utility in small ani-
of the mechanical, electrical and chemical properties mal models, alternative strategies have been designed
of the brain, and the diversity of neural signalling path- for monitoring neural activity at different spatial scales.
ways, may guide materials design to establish intimate Michigan probes are composed of lithographically
long-term synthetic interfaces that are capable of faithful defined patterns of metallic electrodes on thin silicon
recording and stimulation of neural activity. substrates for depth-defined recordings in subcortical
structures26,27. The compatibility of these designs with
History of neural engineering modern complementary metal–oxide–semiconductor
From neuroscience to neural engineering. Motivation processing enables straightforward integration of data
to answer fundamental questions in neuroscience has acquisition and signal amplification capabilities while
consistently inspired the invention of new technologies allowing for facile back-end connectorization27,28. Both
(FIG. 2). In the 1950s, wires were introduced to measure Utah arrays and Michigan probes use the polymers poly-
electrical activity in the nervous system15, followed by imide and parylene C for insulation. Powered by advances
patch-clamp electrophysiology in the 1970s16 to elu- in micro-electro­mechanical systems (MEMS) fabrication,
cidate synaptic transmission. In the 1980s and1990s, the past three decades of neural engineering have deliv-
silicon multielectrode arrays17,18, and stereotrodes and ered a multitude of sophisticated probes inspired by Utah

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 3


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

and Michigan designs with increased resolution, reduced microfabricated devices have to be safeguarded from
dimensions and expanded capabilities29. exposure to warm, aqueous and saline environments,
The emergence of the fields of neuroprosthetics and and oxidative stress associated with the foreign-body
brain–machine interfaces, which seek to restore volun- response. Consequently, strategies for packaging and
tary motor control to patients with paralysis, necessitated encapsulation have emerged as key aspects of neural-
the development of high-resolution recording probes. By probe design. These include hermetic titanium casings
the late 1990s, decoding algorithms had been developed typical of clinical devices, and polyimide, polyurethane,
to control robotic joints using rodent and primate brain poly(dimethyl sulfoxide) (PDMS), SU‑8 and parylene C
activity in real time30,31. The surge of scientific and clinical coatings, which are common in research-grade tools29.
enthusiasm around brain–machine interfaces propelled
this technology into clinical trials in 2004, which pro- Foreign-body response. Neuronal death and the forma-
duced encouraging demonstrations of patients with tetra- tion of glial scars (~100‑μm thick) around implanted
plegia controlling computer cursors25 and robotic arms32 probes can lead to the loss of recorded signals34,36. A sub-
with implanted electrode arrays. However, maintaining ject of several detailed reviews, the glial scar is composed
high SNR recordings from thousands of neurons proved of reactive astrocytes and activated microglia that form
to be challenging, and biocompatibility and reliability dense and electrically inactive tissue along the surface
issues often resulted in probe failure only a few weeks of the device34,36. Several factors have been suggested to
after implantation29,33,34. contribute to this inflammatory response, including:
initial tissue damage during device insertion33,34; elas-
Materials mismatch tic mismatch between the neural probes and the neural
Failure modes of neural probes can be broadly classified tissue in the context of relative micromotion35,37,38; dis-
into those of engineering origin and those of bio­logical ruption of glial networks39; chronic breach of the blood–
origin33–35 (FIG. 3). The former includes mechanical failure brain barrier 40; materials neurotoxicity 41; and chemical
of interconnects between interface boards and implanted mismatch between the implant surface and the cell
probes, degradation of electrical and environmental membranes and extracellular matrix 34.
insulation, and delamination of the different layers Although the initial damage due to device inser-
constituting the device29,35. To avoid premature failure, tion may have short-term effects on the recording

Implantation surgery 1 day 1 week 1 month 1 year

Corrosion

Coating failure

Delamination

Blood–brain barrier breach

Micromotion

Disruption of glial networks Formation of glial scar

Neuronal death

Figure 3 | Mechanisms of neural-probe failure. Failure modes of neural probes, manifested as a loss of neural recording
Nature
capability, can be classified into those related to device design29,35 and those related to foreign-body Reviews
response | Materials
34,36
. Design
failure mechanisms include mechanical failure of interconnects35, degradation and cracking of the insulation217, electrode
corrosion218,219 and delamination of probe layers220. Biological failure mechanisms include initial tissue damage during
insertion33,34; breach of the blood–brain barrier40; elastic mismatch and tissue micromotion35,37,38; disruption of glial
networks39; formation of a glial scar; and neuronal death associated with the above-mentioned factors, as well as with
materials neurotoxicity41 and chemical mismatch34.

4 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

performance, long-term probe performance is of par- strategies to match the mechanics and geometry of the
ticular concern to the brain–machine interface commu- devices to those of neural tissue45.
nity. Introduction of a foreign object with dimensions
exceeding 20 μm has been hypothesized to disrupt local Strategies to improve compatibility
communication between glia, which triggers the release Making stiff electronics soft. Compliant coatings can
of pro-inflammatory cytokines and the recruitment of reduce the apparent elastic mismatch between stiff
additional activated microglia and reactive astrocytes34,39. silicon- or metal-based probes and neural tissue46.
The effects of size are challenging to decouple from those Hydrogels based on synthetic, biological, or a hybrid
associated with elastic mismatch, as smaller device dimen- of these materials can decrease the surface hardness of
sions will also decrease bending stiffness. The elastic mis- the probes, lower electrode impedance and diminish
match between the neural tissue (kilo- to mega­pascals) glial scarring. For example, a poly(vinyl alcohol) matrix
and the implanted probes (1–100 GPa), which are typi- doped with the conductive polymer poly(3,4‑ethylene-
cally tethered to bone, causes repeated injury to the tissue dioxythiophene) (PEDOT) and copolymers of collagen
every time the brain or a nerve is displaced relative to the and the organic semiconductor polypyrrole lowered the
device35,37,38. This continuous impact has been linked to a impedance of lithographically defined metallic elec-
chronic breach of the blood–brain barrier 40, which acti- trodes on silicon or glass substrates while improving cell
vates glia and astrocytes to form a scar around the haem- adhesion47. Incorporation of neuronal adhesion mole­
orrhage. In addition to the mechanical properties of the cules (for example, L1)44, or even live cells48, into the
probe, surface chemistry and topography may influence coating may additionally promote local axonal growth.
the extent of inflammation42. Polymers have been used as alternative substrates
Three decades of neural engineering studies have to support microfabricated probes29,45 (FIG. 4a). Metal
focused on increasing the reliability and biocompatibility electrodes and interconnects can be lithographically
of microfabricated probes through continued miniaturi- defined or microcontact printed onto polyimide and
zation and the addition of coatings to reduce inflamma- parylene C substrates. The same polymers are also used
tion43 and modulus mismatch between the probes and for insulation and packaging. Using microcontact print-
neural tissue46. However, the innate materials properties ing, integrated multiplexed circuits have been integrated
of the implants can limit their biocompatibility, which with flexible polyimide substrates to map the cortical
has motivated the recent emergence of materials-driven dynamics of epileptic seizures with 500‑μm resolution
and millisecond precision49. Conformal chemical vapour
deposition of parylene C onto curved substrates was used
Making stiff electronics soft to create complex sheath and hemispherical cavities with
a Flexible substrates b Wavy, serpentine shapes c Composite fibres internal metal contacts. These flexible neurotrophic elec-
and carbon electrodes trode cavities were filled with growth factors (for exam-
Modulus- ple, nerve growth factor and neurotrophin 3) to promote
matching nerve ingrowth and to establish better interfacial contact
coating
for improved chronic recording performance50,51. PDMS is
also compatible with soft lithography and has been incor-
Polymer porated into devices intended to conform to the dynamic
1–10 μm

Metal
and flexible surfaces of the spinal cord52 and peripheral
Polymer substrate
nerves53–55. Regenerative-sieve55 and nerve-cuff 56 elec-
1–10 μm trodes have also taken advantage of PDMS substrates,
which can be rolled into cylindrical shapes tailored to
Using low-modulus conductive materials the nerve dimensions to support nerve ingrowth.
d Conductive polymers e Microcomposites f Nanocomposites To overcome the challenges associated with buckling
of flexible probes during insertion into the tissue, poly-
mers that undergo a dramatic reduction in modulus from
~1 μm giga- to megapascal in response to temperature (that is,
O O
~<100 nm thermoplastics)57 and hydration (for example, poly(vinyl
S acetate)–cellulose composites inspired by sea-cucumber
S dermis)58 have been explored as stimulus-responsive
n

O O
substrates for neural interfaces.
The bending stiffness of the material (equation 1),
rather than Young’s modulus, determines the tissue
interactions for devices tethered to the skull or vertebrae:
Figure 4 | Examples of approaches intended to overcomeNature Reviews | Materials
the foreign-body
response and increase the resolution of neural interfaces. a | Use of flexible F 48EI
= 3 (1)
substrates and modulus-matching coatings to reduce elastic mismatch between the d L
neural probes and neural tissues. b | Contact-printed serpentine, wavy and fractal
interconnects. c | Thermally drawn fibres with micrometre-scale electrodes and The bending stiffness is defined as the force that is
carbon-fibre microelectrodes. d | Conductive polymers used as flexible alternatives to required to achieve a certain deflection. Here, F is the
crystalline semiconductors and metals. e,f | Conductive composites of flexible and force, d is the deflection, E is Young’s modulus, I is
stretchable polymers and carbon and metal micro- and nanoparticles. the moment of inertia, and L is the length of the device.

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 5


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

For a probe with a rectangular cross-section, inputting the electrode conductivity and the bending stiffness82.
the expression for the moment of inertia yields: A poly­mer matrix can be selected for its low modulus,
and the conductivity, concentration and geometry of
F 4Ewt3
= (2) the micro- or nanoscale dopants determine the elec-
d L3 tronic properties of the blend83 (FIG. 4e,f). Although most
where w and t are the width and thickness of the device, nanocomposites are applied as coatings over metallic
respectively. electrodes, fibre-drawn polyethylene electrodes doped
Because the stiffness scales cubically with the thick- with carbon black were used to record neural activity in
ness of the material (equation 2), reducing the dimen- the spinal cord of a mouse despite having a high imped-
sions of a high-modulus device can dramatically improve ance (>1 MΩ)77 (FIG. 4e). Doping of PDMS with carbon
its flexibility 59,60. This scaling rule has been applied to fab- below the percolation threshold is a promising strategy
ricate compliant microscale wavy surfaces, meshes, ser- for designing microstructured piezo-resistive touch
pentines and fibres composed of materials with Young’s sensors78, and carbon-doped PDMS may eventually be
moduli in the gigapascal range49,61–64 (FIG. 4b,c). Contact integrated into neural recording probes. More recently,
printed meshes and fractal patterns of silicon and metal composites of silver nanowires and styrene–butadiene–
ribbons with thicknesses of a few micrometres can act as styrene elastomers patterned into serpentine meshes with
flexible interconnects between microscale islands bearing elastic moduli of 3–45 kPa (depending on the serpentine
recording electrodes as well as other sensors and process- curvature) and conductivities of ~104 S m−1 were used to
ing electronics49,61. Out‑of‑plane buckling of serpentine modulate cardiac firing in vivo79, illustrating the future
electrodes deposited onto pre-strained PDMS substrates promise of this material platform for functional interfaces
accommodates stretching by up to ~100%, which sig- with other electrically active cells (FIG. 4f). As interest in
nificantly exceeds the strains experienced by peripheral flexible (opto)electronics continues to drive research into
nerves60 (FIG. 4b). Fibre drawing can also be applied to conductive polymer composites, a variety of designs ini-
reduce the dimensions of metallic electrodes embed- tially intended for photovoltaics, OFETs and OLEDs82,84
ded in polymer matrices down to 1–5 μm (REF. 64). Fibre may be integrated into future neural interfaces.
probes with the dimensions of human hair (~80 μm)
can integrate 7–10 electrodes with impedance values Bidirectional neural interfaces
of ~800 kΩ. These probes can record single-unit action Electrical stimulation. Although recording electrodes
potentials and have negligible impact on the surround- provide insight into the electrophysiological activ-
ing neural tissue (FIG. 4c). Similarly, carbon fibre can ity within the nervous system, they do not allow for
be manually separated65,66 or electrospun from carbon control over its dynamics. Electrical stimulation is
nanotube solutions67 to fabricate electrodes that are clinically approved to treat Parkinson disease85 and
5–10 μm in diameter with impedances of 100 kΩ–4 MΩ neuropathic pain86, and is a candidate for alleviating
(FIG. 4c) . The miniature dimensions of the carbon symptoms of major depression87. Chemically stable and
electrodes also permit isolation of individual action bio­compatible platinum and platinum–iridium elec-
potentials and improved tissue compatibility during trodes with low impedance (<10 kΩ) are commonly
chronic implantation65–67. used for electrical stimulation88. The relatively mod-
est charge injection capacity (CIC) of these electrodes
Organic and hybrid electronics at the neural interface. (CIC ≈ 0.05–0.15 mC cm−2) can be improved by coatings
Owing to their low moduli, organic conductors are of iridium oxide (CIC ≈ 5 mC cm−2) or titanium nitride
promising platforms for devices intended to interface (CIC ≈ 0.9 mC cm−2). In addition to reducing electro­
with biological systems68 (FIG. 4d). Aromatic small mol- chemical impedance, PEDOT, polypyrrole and their
ecules and conjugated polymers have been integrated blends with conductive nanomaterials can improve the
into neural probes deployed in vitro69. However, with the CIC (CIC ≈ 1–15 mC cm−2) of the stimulating electrodes88.
exception of PEDOT, the transition of conjugated organ- The CIC can be further improved by increasing the elec-
ics in vivo has been impeded by their poor environmental trode surface area through texturing. Akin to its delete-
stability 70. PEDOT mixed with poly(styrene sulfonate) rious effects on recording electrode performance, glial
(PEDOT:PSS) is a heavily doped p‑type semiconductor scarring diminishes the CIC of stimulating electrodes.
frequently used as a hole-injection and extraction layer Hence, strategies aimed at reducing the inflammatory
in organic light-emitting devices (OLEDs) and photo- response (for example, flexible substrates and modu-
voltaics, and as a channel in organic field-effect tran- lus-matching coatings) may simultaneously improve
sistors (OFETs) or organic electrochemical transistors71. stimulation fidelity 29,45,46,54.
PEDOT:PSS coatings on standard metallic electrodes are
known to increase the SNR and neuronal unit yield72,73. Engineering with optogenetics. Owing to interference
Most recently, flexible PEDOT:PSS organic electrochem- with electrophysiological recording and lack of cellular
ical transistors fabricated on 2‑μm-thick parylene C sub- specificity, the basic mechanisms and therapeutic out-
strates were shown to record local field potentials as well comes of electrical stimulation remain poorly under-
as isolated action potentials from the cortical surface74,75. stood89. Optogenetics can simultaneously perturb and
Blends of polymers and conductive micro- and probe the activity of genetically defined neuronal pop-
nano­particles (such as carbon76–78, metals79,80 and conju- ulations because the optical stimuli are decoupled from
gated organics81) allow for independent control of both electrical recordings90. This method relies on genetic

6 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

introduction of light-sensitive microbial ion channels for multifunctional and flexible optoelectronic probes107
and pumps, called opsins, into mammalian neurons91 (FIG. 5g). Contact transfer printing of metallic inter-
and other electrically active cells92–94. Opsins are trans- connects, electrodes, blue-emitting GaN micro-LEDs,
membrane proteins harbouring the chromophore retinal, microscale thermistors and silicon photodiodes onto
which changes conformation in response to visible light PDMS produced highly compliant multifunctional
and thus drives ion transport 95,96. These proteins can be probes suitable for optogenetic control of mouse behav-
used for both excitation and inhibition of neural activ- iour as well as neural recording, with minimal damage to
ity. Excitatory opsins (for example, channelrhodopsin surrounding tissue108 (FIG. 5h).
2 (ChR2)) are cation channels that mediate membrane Alternatively, polycarbonate and cyclic olefin co­­
depolarization to initiate action potentials in response polymer waveguides and carbon-black–polyethylene
to light. Inhibitory opsins are proton or chloride pumps composite electrodes can be integrated within all-polymer
(for example, halorhodopsin and archaeo­rhodopsin) that probes via high-throughput fibre drawing. The flexible
transport ions against their concentration gradients to fibre probes exhibited a bending stiffness of 10 N m−1
silence neurons through optically driven hyperpolariza- and enabled simultaneous recording and optical neuro-
tion97. Over the past decade, several opsins with varied modulation in the mouse brain and spinal cord without
absorption spectra and photocycle kinetics have been triggering a significant inflammatory reaction64,77 (FIG. 5i).
discovered and engineered98,99.
Because the visible light needed for opsin activation Delivery of chemical stimuli. To perturb neural activity
is scattered and absorbed by neural tissue, optogenetic using traditional pharmacological109 or chemogenetic110
neuromodulation relies on implanted devices to deliver approaches, microfluidic channels can be integrated into
optical stimuli. To take advantage of simultaneous neural probes to deliver chemical and biological agents
electro­physiological recording and optogenetic control, into the nervous system. Optogenetic stimulation aug-
early efforts outfitted mature neural recording tech- mented by delivery of a neuromodulator (for example,
nologies, such as Utah arrays (FIG. 5a), Michigan probes a synaptic blocker) may reduce the ambiguity in opti-
(FIG. 5b), tetrodes and microwires (FIG. 5c), with commer- cal cell-type identification experiments, in which opsin
cially available silica waveguides90,100–102. Multimaterial expression is used to mark a particular neuronal pop-
semiconductor processing methods have recently been ulation102. Furthermore, microfluidic channels permit
applied to monolithic integration of LEDs within silicon injection of viral vectors needed for genetic experiments
probes103,104. For example, indium–gallium–nitride quan- when transgenic animal models are not available111.
tum wells were epitaxially grown on Si(111) substrates Glial scar formation within the probe vicinity may block
to create arrays of up to 12 blue-emitting micro-LEDs the microfluidic channels or change the diffusivity of
(10 μm × 15 μm) integrated with 32 metallic electrodes the compounds of interest. Although integration of
(FIG. 5d) . Sharing geometry with Michigan probes, microfluidic features into flexible MEMS-processed50,107,
these devices enabled depth-defined recording and microcontact-printed111 or fibre-drawn64 neural probes
ChR2‑mediated optogenetic stimulation of pyramidal increases the size of the device, these probes are signifi-
neurons in the rat hippocampus104. An innovative alter- cantly smaller than devices assembled from bulky cannu-
native to silicon probes is a transparent electrode array las, silica optical fibres and metallic electrodes, and hence
composed of the n‑doped (resistivity, ρ ≈ 0.15 Ω∙cm) may reduce the foreign-body response (FIG. 5g–i).
wide-bandgap semiconductor zinc oxide, which per-
mits simultaneous electrical neural recording and deliv- Wireless transmission. Back-end connections scale pro-
ery of optical stimuli with the same material105 (FIG. 5e). portionally with the addition of functional features inte-
Another approach for fabricating transparent probes grated into the implants. The increased device complexity
relies on the integration of graphene electrodes and and weight can affect the behaviour of the experimen-
interconnects onto parylene C substrates. The resulting tal subjects, especially smaller animal models. Wireless
transparent probes permit optical stimulation through power and data transmission can eliminate the need for
the substrates during neural recording 106 (FIG. 5f). multiple tethers and headstages112, but may result in sig-
In addition to bidirectional interface capabilities, nificant heat dissipation within the implanted hardware
biocompatibility should be factored into the design owing to high data rates and power draws. Furthermore,
of integrated optoelectronic probes based on semi­ the physical dimensions of radio­frequency antennas may
conductors. As crystalline group III–V and II–VI not be compatible with implantation into small rodents
semi­conductors exhibit elastic moduli similar to those unless designed for a limited transmission range.
of silicon and metals (tens to hundreds of gigapascals), Recent implants powered by radiofrequency antennas
the tools based on these materials are likely to elicit enable wireless operation of semiconductor micro-LEDs
inflammatory responses similar to those observed for to optically drive behaviour in mice expressing ChR2 in
mature electrode technologies. Consequently, soft- the brain, spinal cord and peripheral nerves113,114. Both
material and hybrid material designs have been explored. direct transmission114 and evanescent coupling through
Using MEMS processing methods, polymer waveguides the animal body 113 have been explored as means to
composed of SU‑8 photoresist were integrated with metal- deliver the necessary power. Direct transmission relied
electrode arrays on polyimide substrates. Although SU‑8 on microcontact serpentine antennas with areas of
can be cytotoxic41 and has limited transmission in the ~3 mm × 3 mm, and the evanescent coupling required
visible spectrum, these initial efforts laid the groundwork a specialized radiofrequency cavity positioned directly

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 7


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Mature electrode arrays outfitted with optical fibres


a b c
ħω

ħω
ħω ħω
ħω

Integrated optoelectronic approaches


d e ħω f
ħω

ħω

ħω
ħω
ħω

Multifunctional neural probes


g ħω h i

ħω
ħω

T
Drug Drug
ħω

Figure 5 | Probes for bidirectional communication with neural circuits. a–c | Integration of optical fibres into Utah
arrays100, Michigan probes101 and tetrodes90, respectively, intended for optogenetics research. d |Nature
Gallium-nitride-based
Reviews | Materials
micro-light-emitting devices (LEDs) monolithically integrated onto silicon substrates within Michigan-style probes104.
e | Transparent multielectrode arrays composed of conductive zinc oxide105. f | Transparent and flexible arrays of graphene
surface electrodes106. g | Micro-electromechanical systems (MEMS)-processed trifunctional polymer probe integrating an
SU‑8 waveguide, metal electrodes and a microfluidic channel on a polyimide substrate107. h | Multifunctional micro­contact-
printed probe containing an electrode, micro-LEDs, a photodetector and a thermistor108. i | Multifunctional all-polymer-fibre
probe integrating a waveguide, carbon composite electrodes and microfluidic channels64. ħω represents the energy of a
photon, with ħ as the reduced Planck constant and ω as the angular frequency; T, temperature in degrees Celcius.

underneath a behavioural enclosure. For applications durations than conventional patch-clamp electro­
requiring additional functionality, such as multiple opti- physiology. Micrometre-sized gold mushrooms, which
cal sources or microfluidic drug delivery, larger antennas were functionalized with RGD (Arg-Gly-Asp) peptides
mounted on the skull were used111. to promote cell adhesion, provided intracellular-like
Independently operating wireless microcircuits that recordings of action potentials as well as subthreshold
are tens to hundreds of micrometres in size, called ‘neural synaptic activity for days in vitro117,118 (FIG. 6a). However,
dust’ motes, have been proposed as an alternative tech­ translation of this technology in vivo is challenging
nology to traditional probes115. Initial millimetre-scale because of its reliance on intimate contact between high-
prototypes have been fabricated and are capable of impedance (~30 MΩ) mushroom electrodes and neu-
recording activity in peripheral nerves116. However, fur- rons. Monocrystalline gold nanowires that are ~100 nm
ther miniaturization poses challenges to device placement in diameter also exhibit high impedances (~5 MΩ),
and stability within the nervous system. but were found to be effective for acute recordings
of single-unit activity in the CA1 region of the mouse
From microprobes to nanotransducers hippo­campus119. Silicon-nanowire and gold-nanopillar
Micro- and nanostructured electrodes permit intra­ electrodes were used for intracellular recordings fol-
cellular recordings and modulation of subthreshold lowing localized membrane electroporation, although
events at finer temporal resolution and over longer intra­cellular access was lost within minutes owing to

8 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

a Endocytosed electrodes b Cell-penetrating nanoprobes c Field-effect devices


Intracellular
Cell Intracellular space Nano-FET
space

Drain
Extracellular
Substrate Substrate Source space

Figure 6 | Micro- and nanoprobes for intracellular recordings. a | Endocytosed mushroom electrodes for high-signal-to‑
Nature Reviews | Materials
noise ratio (SNR) recordings117,118. b | Cell-penetrating nanopillar121, nanowire120 and nanostraw electrodes122.
c | Field-effect devices based on kinked semiconductor nanowires128–130. FET, field-effect transistor.

membrane resealing 120,121. To prolong intracellular access, recording 134 (FIG. 7). Although optical stimulation and
hollow nanotube electrodes that form tight junctions with imaging approaches, such as infrared light135, genetically
the cell membrane were synthesized122. Stealthy probes encoded opsins136,137, and calcium and voltage indica-
with amphiphilic coatings that mimic the chemical struc- tors138, offer superior spatial and temporal resolution,
ture of cell membranes may also provide passive access to they are limited to penetration depths of <1 mm owing
the cytosol through spontaneous fusion123 (FIG. 6b). to tissue absorbance and scattering. Even three-photon
In addition to capacitive voltage recordings, mod- excitation with a 1,675‑nm laser can only access brain
ulation of the gate voltage of semiconductor-nanowire structures that are 1.5-mm deep139.
FETs by the local ion flux can be used for in vitro extra- Acoustic waves can extend the penetration depth by
cellular neural recording 124,125 (FIG. 6c). The same study one to two orders of magnitude in a frequency-dependent
found that nanowire FET arrays are sensitive enough to manner 140,141. Transcranial focused ultrasound at fre-
capture the propagation of action potentials down indi- quencies <0.65 MHz and with an amplitude of 0.12 MPa
vidual axons124. To achieve high-resolution intracellular was shown to modulate activity 3‑cm deep in the human
recordings in three dimensions, FET extension from somatosensory cortex 142. Although ultrasound signals
planar substrates was enabled by the synthesis of kinked with lower frequencies offer deeper tissue penetration
nanowires126,127. Although electrode miniaturization and without significant attenuation, the resolution is propor-
morphological control are promising strategies for long- tional to wavelength, which yields modulation volumes
term intracellular recording, the challenges in device >1 mm3 (REF. 141). Functional ultrasound can map blood
connectorization and implantation currently limit these vessels in the neural tissue based on the characteristic
technologies primarily to in vitro models. Doppler frequency associated with the movement of
To harness the resolution of nanoscale electronics blood cells through the imaging pixels143. Because neural
in vivo, macroporous meshes have been developed that activity within a given region is correlated with increased
have mechanical properties similar to those of neural blood flow, this approach permits non-invasive and
tissue128,129. These metal meshes, supported by SU‑8 sub- indirect measurement of neural dynamics. Without a
strates, served to connect nanowire FETs and platinum contrast agent, a 15‑MHz functional ultrasound probe
nanoelectrodes to back-end electronics and could be enables haemodynamic imaging across the entire rat
injected into the brain. These devices provided chronic brain with a resolution of 100 μm × 100 μm × 200 μm in
single-unit recording while also acting as ‘neurophilic’ 200 ms (REF. 143). Although this method does not rely on
scaffolds that promoted neuronal migration128–130. implanted conduits, it necessitates mounting the func-
Despite the miniaturization and increased flexibility tional ultrasound probe over an aqueous cranial window
of neural probes, a device that eliminates connector- to avoid scattering from the skull.
ized hardware and tissue damage while communicating Magnetic fields are the only modality that can access
with the nervous system over arbitrarily long timescales arbitrarily deep tissues with little attenuation owing to
remains to be demonstrated. By contrast, contactless the low magnetic susceptibility of biological matter 144,145.
stimuli coupled to nanoparticle transducers may facili- Magnetic fields with frequencies <1 kHz and ampli-
tate localized readout or perturbation of neural activity at tudes >1 T can inductively evoke ionic currents to non-
single-protein resolution131. Advances in synthetic meth- invasively modulate brain activity through the skull146.
ods have delivered nanomaterials with tunable size, shape However, transcranial magnetic stimulation is limited
and composition. Furthermore, these nanoparticles can by the coil geometry 147, and clinical applications are con-
be functionalized for cell-specific targeting and improved strained primarily to cortical structures with centimetre
biocompatibility 132,133. spatial resolution. Functional magnetic resonance imag-
ing (fMRI) exploits the differences between the magnetic
Contactless communication properties of oxygenated and deoxygenated haemoglobin
Optical, acoustic and magnetic stimuli offer applica- to measure a local increase in blood flow that accom-
tion-specific advantages for neural stimulation and panies neural activity 148. Although non-invasive, this

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 9


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Skull Brain 10Penetration


–8
depth (m) 10 –9
Spatial resolution (m) Temporal precision (s)

Light
10–6 Light
(skull) 10–7 Light 10–3 Light

Light
(cranial window) 10–4
Light
(brain) 10–5 10–2 Ultrasound
Ultrasound 10 –2
Ultrasound Magnetic
(brain) fields

10–3 Ultrasound 10–1 Magnetic


Ultrasound fields
(cranial window) 100
Magnetic
fields

Magnetic fields 10 10–1 100

Figure 7 | Tissue penetration by optical, ultrasonic and magnetic signals. Electromagnetic waves in the visible and
Nature Reviews | Materials
(near-)infrared optical spectrum afford superior spatial and temporal resolution but limited penetration depth
(~1–1.5 mm) 136,139
. Ultrasound can access deeper brain regions (>50 mm) with a spatial resolution that is inversely
proportional to the wavelength, which, in turn, scales inversely with the penetration depth (in general for ultrasound,
spatial resolution is >1 mm3 and temporal precision is >10 ms)142. Alternating magnetic fields (AMFs) with low
frequencies (<1 kHz) and high amplitudes (0.1–2 T) inductively couple to the upper 1–10 mm of tissue146. AMFs with
amplitudes of ~1–100 mT and frequencies in the low radiofrequency range (0.1–1 MHz) travel through tissue
unaffected145. Temporal precision of neural activity is dependent on the chosen magnetic scheme.

method has limited spatial and temporal resolution also enable the use of these nanocrystals as voltage sen-
with a ~0.5 mm × 0.5 mm voxel area and a sampling rate sors156,157. Technical challenges of using these materials
of <2 Hz. in vivo include the cytotoxicity of quantum dots, which
Nanoscale transducers that convert optical, acoustic are composed primarily of heavy metals (for example,
and magnetic stimuli into biological cues may extend cadmium and lead), and the difficulty in embedding
the signal penetration depth and improve the spatial and them into the cell membrane158. Gold nanoparticles are
temporal resolution of non-invasive neural recording significantly more biocompatible and can couple to light
and modulation approaches131 (FIG. 8). Semiconductor through surface plasmon resonance159. Plasmonic heating
quantum dots, gold nanoparticles and upconversion in gold nanoparticles can induce neural inhibition under
nano­particles can respectively convert light into voltage, constant illumination160 or neural excitation with pulsed
heat or light of a different wavelength (FIG. 8a–c). The laser light at frequencies up to 40 Hz (REF. 161) (FIG. 8b).
deformation induced by ultrasound can be amplified by This photothermal approach allows operation at optical
microbubbles149 or be converted into electric fields by power densities lower than those needed to excite ChR2
piezoelectric materials (FIG. 8d,e). Magnetic nanoparticles while offering subcellular resolution161–163. Because surface
(MNPs) can cluster 150, transduce force151 and apply torque plasmons are sensitive to local electric fields, they may be
in stationary field gradients, or dissipate heat through developed into voltage sensors164, albeit with a narrower
hysteresis in alternating magnetic fields152 (FIG. 8g–i). dynamic range than quantum dots157,165. Nanodiamonds
Interfacing magnetostrictive and ferroelectric materials harbouring nitrogen-vacancy colour centres with photo-
enable conversion of a magnetic field into an electrical luminescence sensitive to temperature166, and electric and
potential153. This list, by no means exhaustive, represents magnetic fields are emerging as biocompatible candidates
a diverse and relatively unexplored materials toolkit for for optical imaging of cellular activity 167,168.
potential next-generation neurotechnologies131. Upconversion nanoparticles absorb near-infrared
light and emit shorter wavelengths through multiphoton
Neural interrogation with nanomaterials processes (FIG. 8c). This effect can extend the penetra-
Optical nanomaterials. Optically active inorganic nano- tion depth of optical stimuli, because tissue scattering is
materials have been explored as non-genetic tools for reduced in the near-infrared-light region169. For example,
precise spatial and temporal control and recording of lanthanide-doped upconversion nanoparticles absorbing
neural activity (FIG. 8a–c). Optical excitation of quantum 980‑nm light were used to control isomerization of the
dots embedded within the cell membrane can, in princi- blue-light-sensitive light-oxygen-voltage (LOV)-sensing
ple, generate sufficient localized electric fields to trigger domain to drive gene transcription170.
action-potential firing 154 (FIG. 8a). However, to date, only
quantum-dot films have been demonstrated to excite Mechanoresponsive nanomaterials. To improve the
neurons upon light exposure154,155. Electric-field-induced resolution of transcranial ultrasound stimulation while
photoluminescence quenching in quantum dots due to maintaining penetration depths of >1 cm (REF. 142),
ionization or the quantum confined Stark effect may mechanoresponsive nanomaterials can be used to

10 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

Optical nanomaterials
a Optoelectronic transitions b Plasmonic heating c Two-photon absorption
ħω ħω Heat
2ħω (infrared)

ħω2 (visible)

Mechanoresponsible nanomaterials
d Micro- and nanobubbles e Piezoelectronic materials f Genetically encoded vesicles


Cavitation
+
Vibration

Magnetic nanomaterials
g MNP clustering h MNP translation or rotation i MNP heating
H
B
H t
∇B

Heat

Intracellular space

Figure 8 | Nanomaterials as local transducers of external stimuli. a | Optoelectronic transitions in semiconductor


Nature Reviews | Materials
quantum dots can be used to convert optical stimuli into electrical signals, and vice versa, which can enable stimulation and
recording of neural activity154–157
. b | Metallic nanoparticles couple to visible and near-infrared light through their plasmon
resonance133. Plasmon energy is dissipated as heat, which can be used to modulate membrane capacitance or trigger
heat-sensitive ion channels in neurons160,161. Plasmon resonance is also sensitive to electric fields and can be used to detect
changes in neuronal membrane voltage164. c | Upconversion nanoparticles convert infrared to visible light through a
two-photon absorption process. This effect may extend the penetration depth of optical neuromodulation techniques170.
d | Vibration and cavitation processes in hollow micro- and nanobubbles may induce action-potential firing in targeted
neurons at ultrasound intensities otherwise insufficient to evoke excitation171. Microbubbles can also improve the resolution
of functional ultrasound imaging by acting as contrast agents149. e | Mechanical deformation by ultrasound can polarize
piezoelectric nanomaterials for localized electrical stimulation172. f | Genetically encoded gas vesicles function similarly to
synthetic inert gas micro- and nanobubbles205. g | Clustering of magnetic nanoparticles (MNPs) in time-constant
magnetic-field gradients can be used to control membrane-protein complexes and influence cell fate150. h | In slowly
varying magnetic fields, magnetic nanodiscs exert torque on the cell membrane178 and MNPs exert tensile force. These
mechanical cues can trigger the opening of mechanosensitive ion channels173. i | Hysteretic heating of MNPs exposed to
alternating magnetic fields triggers heat-sensitive ion channels causing Ca2+ influx and action-potential firing in
neurons181,182. ħω represents the energy of a photon, with ħ as the reduced Planck constant and ω as the angular frequency;
B, magnetic flux density; ∇B, the gradient of the magnetic flux density; H, magnetic field; t, time.

control targeted neurons by local amplification of (FIG. 8d).This effect was linked to the mechanosensi-
low-intensity ultrasound. Perfluorohexane microbub- tive ion channel transient receptor potential 4 (TRP‑4)
bles undergoing physical deformation upon exposure to expressed in sensory neurons, which transduce the
ultrasound (0.2–1 MPa at 2.55 MHz) have been shown localized deformation during microbubble cavitation
to affect the behaviour of Caenorhabditis elegans 171 into a cation influx and firing of action potentials171.

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 11


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

To eliminate the reliance on mechanosensitive chan- Magnetothermal neuromodulation can be formulated


nels, piezoelectric barium titanate (BaTiO3) nano­ as a simple injection into a deep-brain structure sensi-
particles have been engineered to convert ultrasound tized to heat via cell-specific TRPV1 expression to enable
into an electric field172 (FIG. 8e). Ca2+ influx in response chronic stimulation181.
to ultrasound was observed in neuron-like PC12 cells Multiferroic composite nanoparticles consisting
in close proximity to micrometre-sized BaTiO3 aggre- of magnetostrictive and piezoelectric phases enable
gates172. These early examples demonstrate the poten- remote magnetic control of the dipole moment of the
tial of acoustic sensitizers to evoke neural activity when nanoparticle. Magnetoelectric nanomaterials composed
placed near the cell membrane. Future work in primary of magnetostrictive CoFe2O4 nanoparticles coated with
neurons and animal models should assess the efficacy piezoelectric BaTiO3 shells can release electrostatically
of these neuromodulation technologies. In addition adsorbed drugs184, and may be used as transducers for
to their use in neural stimulation, microbubbles and wireless magnetoelectric neuromodulation185.
genetically encoded nanovesicles (FIG. 8e,f) can improve MNPs can also be used as MRI probes of neural activ-
the resolution of functional ultrasound imaging by two ity. At physiological temperatures, isolated superpara-
orders of magnitude, achieving haemodynamic brain magnetic nanoparticles have, on average, zero magnetic
mapping with a pixel size of 8 μm × 10 μm in a given moment. When clustered, the dipole–dipole interac-
illuminated plane149. tions between the nanoparticles impart a non-zero net
magnetic moment onto the aggregates186. The change in
Magnetic nanomaterials. MNPs can transduce a broad magnetic moment can be recorded by MRI as a change
range of stimuli in the presence of magnetic fields and in contrast, allowing for detection of events other than
offer a versatile toolkit for contactless perturbation of cel- haemodynamics187. For example, MNPs conjugated to
lular function (FIG. 8g–i). In a static field gradient, MNPs the Ca2+-binding protein calmodulin enable measure-
transduce force to stimulate the opening of mechano- ment of Ca2+ concentration using MRI188. Although dif-
sensitive ion channels or cluster to switch on aggrega- fusion kinetics currently limit the efficacy of this method
tion-dependent cell pathways173. Relatively large forces in in vivo189, improvement in clustering schemes may sig-
the piconewton range are required to activate mechano­ nificantly enhance sensor sensitivity and the dynamic
sensitive ion channels174, which necessitates the use of range190. With further refinement, these nanoparticle
high field gradients (>5 T m−1) and particles with large sensors may offer higher signal contrast at lower con-
magnetic moments to evoke a cellular event 151,173,175,176. centrations than current small-molecule agents used for
By contrast, forces of only a few femtonewtons can drive real-time monitoring of neurotransmitter levels in the
lateral association of receptors embedded in lipid mem- brain using MRI191.
branes (FIG. 8g). This mechanism can induce apoptosis
through clustering of MNPs linked to death receptor 4 Prospects and challenges
(also known as TRAILR1 and TNFRSF10A) in an applied Emerging applications of nanomaterials at the neural
field of 0.2 T (REF. 150). Because the magnetic moment and interface demonstrate the possibility of contactless neuro­
targeting moieties of MNPs can be varied independently modulation and recording without fixed hardware and
through their size, composition and surface functional- transgenes. Further gain in function can be envisioned
ization, various mechanical and chemical cues can be through combinatorial strategies involving multiple nano-
delivered with high spatial and temporal precision177. materials and external stimuli. For example, independent
Alternating magnetic fields (AMFs) can also be neuromodulation of different cell populations may be
applied to control cellular function. For example, possible by tailoring the nanomaterial properties, such as
micrometre-sized magnetizable ferrite-doped beads using MNPs that selectively dissipate heat in distinct AMF
and permalloy discs can disrupt the cytoskeleton or conditions or plasmonic gold nanorods that absorb light
cell membrane by exerting torque in a slowly varying at different wavelengths192,193. Materials optimization for a
field (10–50 Hz)178,179 (FIG. 8h). AMFs with frequencies of specific transduction mechanism can reduce stimulation
up to 5 kHz coupled to superparamagnetic zinc-doped latency 194 or enhance reporter sensitivity 157. Finally, inte-
ferrite (Zn0.4Fe2.6O4) nanocubes can induce deflection gration of probes that transduce orthogonal stimuli may
and Ca2+ influx in mechanosensitive inner-ear hair further expand the experimental palette.
cells151 (FIG. 8h). In AMFs with low radiofrequencies The delivery of nanomaterials in vivo is perhaps a big-
(100 kHz−1 MHz) and amplitudes (10–100 mT), chem- ger design challenge than optimizing the intrinsic prop-
ically inert ferrite MNPs (MexFe3 − xO4, Me = Fe, Co, erties of nanomaterials195. The ability of nanoparticles
Mn or Zn) can undergo hysteresis and dissipate heat 180 to target the site of interest and their residential lifetime
(FIG. 8i). This phenomenon, which arises from irrevers- at the neural interface depends on a number of factors,
ible work done during a magnetization cycle, has been including the surface chemistry, size, shape, composi-
investigated as a potential cancer treatment for nearly tion and concentration of the nanoparticles. Although
five decades. More recently, magnetic hyperthermia antibody conjugation to nanoparticles affords speci-
was applied to the remote control of action-potential ficity in vitro, nonspecific protein adsorption and poor
firing 181,182 and gene transcription 183. AMF-driven colloidal stability may impede this targeting approach
heat dissipation by MNPs can trigger the opening of in vivo 196,197. Difficulty in traversing the blood–brain
the heat-sensitive channel TRPV1, which can lead to barrier 195 and transient lifetime on the cell membrane198
Ca2+ influx and action-potential firing in neurons181,182. render cell-specific and chronic neural interfaces with

12 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

nanoparticles technically challenging. Non-surgical strat- Conclusion and outlook


egies to deliver nanoparticles to a particular brain region Multifunctional devices capable of forming intimate
include temporary breach of the blood–brain barrier interfaces with neurons without provoking a severe
using focused ultrasound199 and, for MNPs, spatial local- foreign-body response would provide detailed maps of
ization with magnetic-field gradients200. Improvement of neural activity and serve as neuroprosthetic links with
these delivery methods may eventually allow nanoparti- electronic circuits. Although nano- and microfabricated
cles to target a particular neural circuit by periodic intra- neural probes offer unprecedented resolution and tempo-
venous or oral administration to overcome the transient ral precision, the mismatch in mechanical and chemical
nature of the nanomaterial–neuron interface. When inva- properties of metals, oxides and semiconductors with
sive delivery is acceptable, nanoparticles can be directly neural tissue has frequently led to probe failure with time.
injected into the region of interest. Amino-functionalized Fuelled by advances in materials synthesis and flex-
MNPs can persist for up to a year in the human ible electronics, materials-centric designs have steadily
brain201. In the peripheral nervous system, nanoparti- reduced the foreign-body response at the brain–machine
cles can be injected into a targeted nerve encapsulated interface over the past decade. Concomitantly, the devel-
within a scaffolding material that delays nanoparticle opment of opto- and chemogenetic tools has brought
clearance202–204. new opportunities for neuroscientists and inspired engi-
Biogenic nanomaterials can circumvent the diffi- neers to integrate optical and microfluidic features into
culties in cell-specific nanoparticle targeting and their neural probes. However, the reliance on fixed hardware
clearance through constant expression of the relevant for signal transduction still limits the number of neu-
protein complexes from the integrated transgenes. Gas- rons that can be independently interrogated across the
filled vesicles can be genetically encoded and engineered nervous system.
to improve temporal and spatial resolution of functional Synthetic nanomaterials, with their diverse and
ultrasound imaging 205. Magnetogenetic excitation tunable properties, can transduce optical, acoustic
and inhibition of neural activity in mice was recently and magnetic stimuli into biological signals, and vice
reported with ion channels fused to ferritin206,207, a pro- versa. Although targeting of nanomaterials into specific
tein that can mineralize iron into weakly paramagnetic locations of the nervous system presents a formidable
ferrihydrite208. Unlike synthetic MNPs composed of fer- challenge, the potential payoff is the ability to remotely
rimagnetic iron oxide and its alloys, ferritin possesses a control and record the activity of billions of neurons.
negligible magnetic moment, and the biophysical ori- Genetically encoded reporters and mediators of neural
gins of the observed physiological response to applied activity have improved our understanding of brain and
magnetic fields thus remain unclear 209. Magnetotactic nerve function, and have provided insight into future
bacteria are known to produce magnetite nanoparticles, therapies for neurological disorders. Their clinical trans-
which have significantly higher magnetic moments than lation, however, remains uncertain and will require new
ferritin210. However, transferring their genetic machinery technologies that will probably operate at the nanoscale
into mammalian cells requires further optimization211. to conform to the complexity of the nervous system.

1. Dorsey, E. R. et al. Projected number of people with 12. Patapoutian, A. et al. ThermoTRP channels and 19. McNaughton, B. L., O’Keefe, J. & Barnes, C. A. The
Parkinson disease in the most populous nations, 2005 beyond: mechanisms of temperature sensation. Nat. stereotrode: a new technique for simultaneous
through 2030. Neurology 68, 384–386 Rev. Neurosci. 4, 529–539 (2003). isolation of several single units in the central nervous
(2007). 13. Jordt, S. E., McKemy, D. D. & Julius, D. Lessons from system from multiple unit records. J. Neurosci.
2. Adelman, G., Rane, S. G. & Villa, K. F. The cost burden peppers and peppermint: the molecular logic of Methods 8, 391–397 (1983).
of multiple sclerosis in the United States: a systematic thermosensation. Curr. Opin. Neurobiol. 13, 20. Gray, C. M., Maldonado, P. E., Wilson, M. &
review of the literature. J. Med. Econ. 16, 639–647 487–492 (2003). McNaughton, B. Tetrodes markedly improve the
(2013). 14. Delmas, P., Hao, J. & Rodat-Despoix, L. Molecular mech­ reliability and yield of multiple single-unit isolation
3. Greenberg, P. E. et al. The economic burden of anisms of mechanotransduction in mammalian sensory from multi-unit recordings in cat striate cortex.
depression in the United States: how did it change neurons. Nat. Rev. Neurosci. 12, 139–153 (2011). J. Neurosci. Methods 63, 43–54 (1995).
between 1990 and 2000? J. Clin. Psychiatry 64, 15. Strumwasser, F. Long-term recording from single Combining four microwires into a tetrode
1465–1475 (2003). neurons in brain of unrestrained mammals. Science arrangement was shown to yield superior
4. Mathers, C. D. & Loncar, D. Projections of global 127, 469–470 (1958). identification of isolated action potentials.
mortality and burden of disease from 2002 to 2030. This article is the first report of a chronic recording 21. Fee, M. S. & Leonardo, A. Miniature motorized
PLoS Med. 3, e442 (2006). of isolated action potentials in the brain of a freely microdrive and commutator system for chronic neural
5. Tracey, K. J. Reflex control of immunity. Nat. Rev. moving mammal (a ground squirrel). recording in small animals. J. Neurosci. Methods 112,
Immunol. 9, 418–428 (2009). 16. Sakmann, B. & Neher, E. Patch clamp techniques for 83–84 (2001).
6. Kandel, E. R., Schwartz, J. H. & Jessell, T. M. Principles studying ionic channels in excitable membranes. Annu. 22. Herry, C. et al. Switching on and off fear by distinct
of Neural Science 4th edn (McGraw-Hill Medical, 2000). Rev. Physiol. 46, 455–472 (1984). neuronal circuits. Nature 454, 600–606
7. Azevedo, F. A. et al. Equal numbers of neuronal and 17. Campbell, P. K., Jones, K. E., Huber, R. J., Horch, K. W. (2008).
nonneuronal cells make the human brain an & Normann, R. A. A silicon-based, three-dimensional 23. Wilson, M. A. & McNaughton, B. L. Reactivation of
isometrically scaled-up primate brain. J. Comp. Neurol. neural interface: manufacturing processes for an hippocampal ensemble memories during sleep.
513, 532–541 (2009). intracortical electrode array. IEEE Trans. Biomed. Eng. Science 265, 676–679 (1994).
8. Haydon, P. G. Glia: listening and talking to the 38, 758–768 (1991). 24. Bartels, J. et al. Neurotrophic electrode: method of
synapse. Nat. Rev. Neurosci. 2, 185–193 (2001). This article presents an early conceptual assembly and implantation into human motor
9. Barres, B. A. The mystery and magic of glia: a demonstration of the Utah array. speech cortex. J. Neurosci. Methods 174, 168–176
perspective on their roles in health and disease. Neuron 18. Drake, K. L., Wise, K. D., Farraye, J., Anderson, D. J. & (2008).
60, 430–440 (2008). BeMent, S. L. Performance of planar multisite 25. Hochberg, L. R. et al. Neuronal ensemble control of
10. Pakkenberg, B. et al. Aging and the human neocortex. microprobes in recording extracellular single-unit prosthetic devices by a human with tetraplegia.
Exp. Gerontol. 38, 95–99 (2003). intracortical activity. IEEE Trans. Biomed. Eng. 35, Nature 442, 164–171 (2006).
11. Wemmie, J. A., Taugher, R. J. & Kreple, C. J. Acid- 719–732 (1988). This article demonstrates the use of neural signals
sensing ion channels in pain and disease. Nat. Rev. In this article, Michigan probes were applied to from the brain of a patient with tetraplegia to
Neurosci. 14, 461–471 (2013). record neural activity. control a prosthetic arm.

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 13


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

26. Buzsaki, G. et al. High-frequency network oscillation in microfabricated PDMS microchannel electrode array. elastomeric fibres. Nat. Nanotechnol. 7, 803–809
the hippocampus. Science 256, 1025–1027 (1992). Lab Chip 12, 2540–2551 (2012). (2012).
27. Wise, K. D. Silicon microsystems for neuroscience and 54. Lacour, S. P. et al. Flexible and stretchable micro- 81. Abidian, M. R., Corey, J. M., Kipke, D. R. &
neural prostheses. IEEE Eng. Med. Biol. Mag. 24, electrodes for in vitro and in vivo neural interfaces. Martin, D. C. Conducting-polymer nanotubes improve
22–29 (2005). Med. Biol. Eng. Comput. 48, 945–954 (2010). electrical properties, mechanical adhesion, neural
28. Yazicioglu, F. et al. Ultra-high-density in vivo neural 55. Srinivasan, A. et al. Microchannel-based regenerative attachment, and neurite outgrowth of neural
probes. Conf. Proc. IEEE Eng. Med. Biol. Soc. scaffold for chronic peripheral nerve interfacing in electrodes. Small 6, 421–429 (2010).
2032–2035 (2014). amputees. Biomaterials 41, 151–165 (2015). 82. Yao, S. & Zhu, Y. Nanomaterial-enabled stretchable
29. Scholten, K. & Meng, E. Materials for microfabricated 56. FitzGerald, J. J. et al. A regenerative microchannel conductors: strategies, materials and devices. Adv.
implantable devices: a review. Lab Chip 15, neural interface for recording from and stimulating Mater. 27, 1480–1511 (2015).
4256–4272 (2015). peripheral axons in vivo. J. Neural Eng. 9, 016010 83. Kirkpatrick, S. Percolation and conduction. Rev. Mod.
30. Nicolelis, M. A. L. Brain–machine interfaces to restore (2012). Phys. 45, 574–588 (1973).
motor function and probe neural circuits. Nat. Rev. 57. Ware, T. et al. Fabrication of responsive, softening 84. Hu, L. et al. Scalable coating and properties of
Neurosci. 4, 417–422 (2003). neural interfaces. Adv. Funct. Mater. 22, 3470–3479 transparent, flexible, silver nanowire electrodes. ACS
31. Donoghue, J. P. Connecting cortex to machines: (2012). Nano 4, 2955–2963 (2010).
recent advances in brain interfaces. Nat. Neurosci. 5, 58. Capadona, J. R., Shanmuganathan, K., Tyler, D. J. & 85. Perlmutter, J. S. & Mink, J. W. Deep brain stimulation.
1085–1088 (2002). Rowan, S. J. Weder, C. Stimuli-responsive polymer Annu. Rev. Neurosci. 29, 229–257 (2006).
32. Hochberg, L. R. et al. Reach and grasp by people with nanocomposites inspired by the sea cucumber dermis. 86. North, R. B. et al. Spinal cord stimulation for chronic,
tetraplegia using a neurally controlled robotic arm. Science 319, 1370–1374 (2008). intractable pain: superiority of “multi-channel” devices.
Nature 485, 372–375 (2012). 59. Kim, D. H. et al. Flexible and stretchable electronics Pain 44, 119–130 (1991).
33. Ward, M. P., Rajdev, P., Ellison, C. & Irazoqui, P. P. for biointegrated devices. Annu. Rev. Biomed. Eng. 14, 87. Holtzheimer, P. E. III & Mayberg, H. S. Deep brain
Toward a comparison of microelectrodes for acute and 113–128 (2012). stimulation for treatment-resistant depression. Am.
chronic recordings. Brain Res. 1282, 183–200 (2009). 60. Rogers, J. A., Someya, T. & Huang, Y. Materials and J. Psychiatry 167, 1437–1444 (2010).
34. Polikov, V. S., Tresco, P. A. & Reichert, W. M. Response mechanics for stretchable electronics. Science 327, 88. Cogan, S. F. Neural stimulation and recording
of brain tissue to chronically implanted neural 1603–1607 (2010). electrodes. Annu. Rev. Biomed. Eng. 10, 275–309
electrodes. J. Neurosci. Methods 148, 1–18 (2005). 61. Kim, D. H. et al. Epidermal electronics. Science 333, (2008).
A comprehensive review article that summarized 838–343 (2011). 89. Kringelbach, M. L., Jenkinson, N., Owen, S. L. F. &
the biological failure modes seen in chronically A pioneering application of flexible and stretchable Aziz, T. Z. Translational principles of deep brain
implanted neural probes. microcontact-printed electronics for biological stimulation. Nat. Rev. Neurosci. 8, 623–635 (2007).
35. James, C. B. et al. Failure mode analysis of silicon- sensing. 90. Anikeeva, P. et al. Optetrode: a multichannel readout
based intracortical microelectrode arrays in non- 62. Kim, R. H. et al. Waterproof AlInGaP optoelectronics for optogenetic control in freely moving mice. Nat.
human primates. J. Neural Eng. 10, 066014 (2013). on stretchable substrates with applications in Neurosci. 15, 163–170 (2011).
36. Kozai, T. D. Y. et al. Brain tissue responses to neural biomedicine and robotics. Nat. Mater. 9, 929–937 91. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G., &
implants impact signal sensitivity and intervention (2010). Deisseroth, K. Millisecond-timescale, genetically
strategies. ACS Chem. Neurosci. 6, 48–67 (2015). 63. Fan, J. A. et al. Fractal design concepts for stretchable targeted optical control of neural activity. Nat.
37. Lee, H., Bellamkonda, R. V., Sun, W. & electronics. Nat. Commun. 5, 3266 (2014). Neurosci. 8, 1263–1268 (2005).
Levenston, M. E. Biomechanical analysis of silicon 64. Canales, A. et al. Multifunctional fibers for ChR2, a microbial opsin, is used to control neural
microelectrode-induced strain in the brain. J. Neural simultaneous optical, electrical and chemical activity with millisecond optical pulses for the first
Eng. 2, 81–89 (2005). interrogation of neural circuits in vivo. Nat. Biotechnol. time, marking the invention of optogenetics.
38. Lind, G., Linsmeier, C. E. & Schouenborg, J. The 33, 277–284 (2015). 92. Bruegmann, T. et al. Optogenetic control of heart muscle
density difference between tissue and neural probes is 65. Kozai, T. D. et al. Ultrasmall implantable composite in vitro and in vivo. Nat. Methods 7, 897–900 (2010).
a key factor for glial scarring. Sci. Rep. 3, 2942 (2013). microelectrodes with bioactive surfaces for chronic 93. Magown, P. et al. Direct optical activation of skeletal
39. Szarowski, D. H. et al. Brain responses to micro- neural interfaces. Nat. Mater. 11, 1065–1073 muscle fibres efficiently controls muscle contraction
machined silicon devices. Brain Res. 983, 23–35 (2012). and attenuates denervation atrophy. Nat. Commun. 6,
(2003). 66. Guitchounts, G., Markowitz, J. E., Liberti, W. A. & 8506 (2015).
40. Saxena, T. et al. The impact of chronic blood–brain Gardner, T. J. A carbon-fiber electrode array for long- 94. Montgomery, K. L. et al. Beyond the brain:
barrier breach on intracortical electrode function. term neural recording. J. Neural Eng. 10, 046016 optogenetic control in the spinal cord and peripheral
Biomaterials 34, 4703–4713 (2013). (2013). nervous system. Sci. Transl. Med. 8, 337rv5 (2016).
41. Kotzar, G. et al. Evaluation of MEMS materials of 67. Vitale, F. et al. Neural stimulation and recording with 95. Nagel, G. et al. Channelrhodopsin 2, a directly light-
construction for implantable medical devices. bidirectional, soft carbon nanotube fiber gated cation-selective membrane channel. Proc. Natl
Biomaterials 23, 2737–2750 (2002). microelectrodes. ACS Nano 9, 4465–4474 (2015). Acad. Sci. USA 100, 13940–13945 (2003).
42. Chapman, C. A. R. et al. Nanoporous gold as a neural 68. Malliaras, G. G. Organic bioelectronics: a new era for 96. Zhang, F. et al. The microbial opsin family of
interface coating: effects of topography, surface organic electronics. Biochim. Biophys. Acta 1830, optogenetic tools. Cell 147, 1446–1457 (2011).
chemistry, and feature size. ACS Appl. Mater. 4286–4287 (2013). 97. Zhang, F. et al. Multimodal fast optical interrogation of
Interfaces 7, 7093–7100 (2015). 69. Benfenati, V. et al. A transparent organic transistor neural circuitry. Nature 446, 633–639 (2007).
43. Zhong, Y. & Bellamkonda, R. V. Dexamethasone- structure for bidirectional stimulation and recording of 98. Yizhar, O., Fenno, L. E., Davidson, T. J., Mogri, M. &
coated neural probes elicit attenuated inflammatory primary neurons. Nat. Mater. 12, 672–680 (2013). Deisseroth, K. Optogenetics in neural systems. Neuron
response and neuronal loss compared to uncoated 70. Kawano, K. et al. Degradation of organic solar cells 71, 9–34 (2011).
neural probes. Brain Res. 1148, 15–27 (2007). due to air exposure. Sol. Energy Mater. Sol. Cells 90, 99. Klapoetke, N. C. et al. Independent optical excitation
44. Azemi, E. et al. Surface immobilization of neural 3520–3530 (2006). of distinct neural populations. Nat. Methods 11,
adhesion molecule L1 for improving the 71. Elschner, A., Kirchmeyer, S., Lovenich, W., Merker, U. 338–346 (2014).
biocompatibility of chronic neural probes: in vitro & Reuter, K. PEDOT: Principles and Applications of an 100. Zhang, J. et al. Integrated device for optical
characterization. Acta Biomater. 4, 1208–1217 Intrinsically Conductive Polymer (CRC Press, 2010). stimulation and spatiotemporal electrical recording of
(2008). 72. Ludwig, K. A. et al. Chronic neural recordings using neural activity in light-sensitized brain tissue. J. Neural
45. Jeong, J. W. et al. Soft materials in neuroengineering silicon microelectrode arrays electrochemically Eng. 6, 055007 (2009).
for hard problems in neuroscience. Neuron 86, deposited with a poly(3,4 ethylenedioxythiophene) 101. Royer, S. et al. Multi-array silicon probes with
175–186 (2015). (PEDOT) film. J. Neural Eng. 3, 59–70 (2006). integrated optical fibers: light-assisted perturbation
46. Aregueta-Robles, U. A. et al. Organic electrode 73. Ludwig, K. A. et al. Poly(3,4 ethylenedioxythiophene) and recording of local neural circuits in the behaving
coatings for next-generation neural interfaces. Front. (PEDOT) polymer coatings facilitate smaller neural animal. Eur. J. Neurosci. 31, 2279–2291 (2010).
Neuroeng. 7, 15 (2014). recording electrodes. J. Neural Eng. 8, 014001 102. Kravitz, A. V., Owen, S. F. & Kreitzer, A. C. Optogenetic
47. Kim, D. H., Wiler, J. A., Anderson, D. J., Kipke, D. R. & (2011). identification of striatal projection neuron subtypes
Martin, D. C. Conducting polymers on hydrogel-coated 74. Khodagholy, D. et al. In vivo recordings of brain during in vivo recordings. Brain Res. 1511, 21–32
neural electrode provide sensitive neural recordings in activity using organic transistors. Nat. Commun. 4, (2013).
auditory cortex. Acta Biomater. 6, 57–62 (2010). 1575 (2013). 103. Buzsáki, G. et al. Tools for probing local circuits: high-
48. Green, R. A., Lovell, N. H. & Poole-Warren, L. A. Cell 75. Khodagholy, D. et al. NeuroGrid: recording action density silicon probes combined with optogenetics.
attachment functionality of bioactive conducting potentials from the surface of the brain. Nat. Neuron 86, 92–105 (2015).
polymers for neural interfaces. Biomaterials 30, Neurosci. 18, 310–315 (2015). 104. Wu, F. et al. Monolithically integrated μLEDs on silicon
3637–3644 (2009). 76. Castagnola, E. et al. Biologically compatible neural neural probes for high-resolution optogenetic studies
49. Viventi, J. et al. Flexible, foldable, actively multiplexed, interface to safely couple nanocoated electrodes to in behaving animals. Neuron 88, 1136–1148 (2015).
high-density electrode array for mapping brain activity the surface of the brain. ACS Nano 7, 3887–3895 105. Lee, J., Ozden, I., Song, Y. K. & Nurmikko, A. V.
in vivo. Nat. Neurosci. 14, 1599–1605 (2011). (2013). Transparent intracortical microprobe array for
50. Kim, B. J. et al. 3D Parylene sheath neural probe for 77. Lu, C. et al. Polymer fiber probes enable optical simultaneous spatiotemporal optical stimulation and
chronic recordings. J. Neural Eng. 10, 045002 (2013). control of spinal cord and muscle function in vivo. Adv. multichannel electrical recording. Nat. Methods 12,
51. Kuo, J. T. W. et al. Novel flexible Parylene neural probe Funct. Mater. 24, 6594–6600 (2014). 1157–1162 (2015).
with 3D sheath structure for enhancing tissue 78. Tee, B. C. K. et al. A skin-inspired organic digital 106. Kuzum, D. et al. Transparent and flexible low noise
integration. Lab Chip 13, 554–561 (2012). mechanoreceptor. Science 350, 313–316 (2015). graphene electrodes for simultaneous electrophysiology
52. Minev, I. R. et al. Electronic dura mater for long-term 79. Park, J. et al. Electromechanical cardioplasty using a and neuroimaging. Nat. Commun. 5, 5259 (2014).
multimodal neural interfaces. Science 347, 159–163 wrapped elasto-conductive epicardial mesh. Sci. 107. Rubehn, B., Wolff, S. B., Tovote, P., Lüthi, A. &
(2015). Transl. Med. 8, 344ra86 (2016). Stieglitz, T. A polymer-based neural microimplant for
53. Delivopoulos, E. et al. Concurrent recordings of 80. Park, M. et al. Highly stretchable electric circuits from optogenetic applications: design and first in vivo study.
bladder afferents from multiple nerves using a a composite material of silver nanoparticles and Lab Chip 13, 579–588 (2013).

14 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

108. Kim, T. I. et al. Injectable, cellular-scale optoelectronics 134. Luan, S. et al. Neuromodulation: present and 161. Carvalho-de Souza, J. L. et al. Photosensitivity of
with applications for wireless optogenetics. Science emerging methods. Front. Neuroeng. 7, 27 (2014). neurons enabled by cell-targeted gold nanoparticles.
340, 211–216 (2013). 135. Wells, J. et al. Application of infrared light for in vivo Neuron 86, 207–217 (2015).
109. van den Brand, R. et al. Restoring voluntary control of neural stimulation. J. Biomed. Opt. 10, 064003 (2005). 162. Lavoie-Cardinal, F. et al. Gold nanoparticle-assisted all
locomotion after paralyzing spinal cord injury. Science 136. Yizhar, O. et al. Optogenetics in neural systems. optical localized stimulation and monitoring of Ca2+
336, 1182–1185 (2012). Neuron 71, 9–34 (2011). signaling in neurons. Sci. Rep. 6, 20619 (2016).
110. Urban, D. J. & Roth, B. L. DREADDs (designer 137. Aravanis, A. M. et al. An optical neural interface: in 163. Nakatsuji, H. et al. Thermosensitive ion channel
receptors exclusively activated by designer drugs): vivo control of rodent motor cortex with integrated activation in single neuronal cells by using surface-
chemogenetic tools with therapeutic utility. Annu. Rev. fiberoptic and optogenetic technology. J. Neural Eng. engineered plasmonic nanoparticles. Angew. Chem.
Pharmacol. Toxicol. 55, 399–417 (2015). 4, S143–S156 (2007). Int. Ed. 54, 11725–11729 (2015).
This comprehensive review describes powerful 138. Knopfel, T. Genetically encoded optical indicators for 164. Zhang, J., Atay, T. & Nurmikko, A. V. Optical detection
chemogenetic approaches to targeted the analysis of neuronal circuits. Nat. Rev. Neurosci. of brain cell activity using plasmonic gold
neuromodulation with DREADDs (designer 13, 687–700 (2012). nanoparticles. Nano Lett. 9, 519–524 (2009).
receptors exclusively activated by designer drugs). 139. Horton, N. G. et al. In vivo three-photon microscopy of 165. Juluri, B. K. et al. Effects of geometry and composition
111. Jeong, J. W. et al. Wireless optofluidic systems for subcortical structures within an intact mouse brain. on charge-induced plasmonic shifts in gold
programmable in vivo pharmacology and Nat. Photonics 7, 205–209 (2013). nanoparticles. J. Phys. Chem. C. 112, 7309–7317
optogenetics. Cell 162, 662–674 (2015). 140. Han, F.-F. & Hu, J.-H. in SPAWDA 2015, Symp. (2008).
112. Muller, R. et al. A minimally invasive 64 channel Piezoelectricity Acoust. Waves Device Appl. 102–105 166. Kucsko, G. et al. Nanometre-scale thermometry in a
wireless μECoG implant. IEEE J. Solid-State Circuits (SPAWDA, 2015). living cell. Nature 500, 54–58 (2013).
50, 344–359 (2015). 141. Bystritsky, A. & Korb, A. A review of low-intensity 167. Karaveli, S. et al. Modulation of nitrogen vacancy
113. Montgomery, K. L. et al. Wirelessly powered, fully transcranial focused ultrasound for clinical charge state and fluorescence in nanodiamonds using
internal optogenetics for brain, spinal and peripheral applications. Curr. Behav. Neurosci. Rep. 2, 60–66 electrochemical potential. Proc. Natl Acad. Sci. USA
circuits in mice. Nat. Methods 12, 969–974 (2015). 113, 3938–3943 (2016).
(2015). 142. Legon, W. et al. Transcranial focused ultrasound 168. Barry, J. et al. Optical magnetic detection of single-
114. Park, S. I. et al. Soft, stretchable, fully implantable modulates the activity of primary somatosensory cortex neuron action potentials using quantum defects in
miniaturized optoelectronic systems for wireless in humans. Nat. Neurosci. 17, 322–329 (2014). diamond. Preprint at https://arxiv.org/abs/1602.01056
optogenetics. Nat. Biotechnol. 33, 1280–1286 Focused ultrasound is used for non-invasive (2016).
(2015). neuromodulation in the cortex of human subjects. 169. Zhang, Y. et al. Illuminating cell signaling with near-
115. Seo, D. et al. Model validation of untethered, 143. Mace, E. et al. Functional ultrasound imaging of the infrared light-responsive nanomaterials. ACS Nano.
ultrasonic neural dust motes for cortical recording. brain. Nat. Methods 8, 662–664 (2011). 10, 3881–3885 (2016).
J. Neurosci. Methods 244, 114–122 (2015). 144. Bottomley, P. A. & Andrew, E. R. RF magnetic field 170. He, L. et al. Near-infrared photoactivatable control of
116. Seo, D. et al. Wireless recording in the peripheral penetration, phase shift and power dissipation in Ca2+ signaling and optogenetic immunomodulation.
nervous system with ultrasonic neural dust. Neuron biological tissue: implications for NMR imaging. Phys. eLife 4, e10024 (2015).
91, 529–539 (2016). Med. Biol. 23, 630–643 (1978). 171. Ibsen, S. et al. Sonogenetics is a non-invasive
The first in vivo validation of neural dust motes. 145. Young, J. H., Wang, M. T. & Brezovich, I. A. Frequency/ approach to activating neurons in Caenorhabditis
117. Spira, M. E. & Hai, A. Multi-electrode array depth-penetration considerations in hyperthermia by elegans. Nat. Commun. 6, 8264 (2015).
technologies for neuroscience and cardiology. Nat. magnetically induced currents. Electron. Lett. 16, 172. Marino, A. et al. Piezoelectric nanoparticle-assisted
Nanotechnol. 8, 83–94 (2013). 358–359 (1980). wireless neuronal stimulation. ACS Nano 9,
118. Shmoel, N. et al. Multisite electrophysiological 146. Hallett, M. Transcranial magnetic stimulation: a 7678–7689 (2015).
recordings by self-assembled loose-patch-like junctions primer. Neuron 55, 187–199 (2007). 173. Dobson, J. Remote control of cellular behaviour with
between cultured hippocampal neurons and 147. Deng, Z. D., Lisanby, S. H. & Peterchev, A. V. Electric magnetic nanoparticles. Nat. Nanotechnol. 3,
mushroom-shaped microelectrodes. Sci. Rep. 6, 27110 field depth–focality tradeoff in transcranial magnetic 139–143 (2008).
(2016). stimulation: simulation comparison of 50 coil designs. 174. Ingber, D. E. Cellular mechanotransduction: putting all
119. Kang, M. et al. Subcellular neural probes from single- Brain Stimul. 6, 1–13 (2013). the pieces together again. FASEB. J. 20, 811–827
crystal gold nanowires. ACS Nano 8, 8182–8189 148. Matthews, P. M., Honey, G. D. & Bullmore, E. T. (2006).
(2014). Applications of fMRI in translational medicine and 175. Seo, D. et al. A mechanogenetic toolkit for
120. Robinson, J. T. et al. Vertical nanowire electrode arrays clinical practice. Nat. Rev. Neurosci. 7, 732–744 interrogating cell signaling in space and time. Cell
as a scalable platform for intracellular interfacing to (2006). 165, 1507–1518 (2016).
neuronal circuits. Nat. Nanotechnol. 7, 180–184 149. Errico, C. et al. Ultrafast ultrasound localization 176. Hughes, S. et al. Selective activation of
(2012). microscopy for deep super-resolution vascular mechanosensitive ion channels using magnetic
121. Xie, C. et al. Intracellular recording of action potentials imaging. Nature 527, 499–502 (2015). particles. J. R. Soc. Interface 5, 855–863 (2008).
by nanopillar electroporation. Nat. Nanotechnol. 7, 150. Cho, M. H. et al. A magnetic switch for the control of 177. Hoffmann, C. et al. Spatiotemporal control of
185–190 (2012). cell death signalling in in vitro and in vivo systems. microtubule nucleation and assembly using magnetic
122. Lin, Z. C. et al. Iridium oxide nanotube electrodes for Nat. Mater. 11, 1038–1043 (2012). nanoparticles. Nat. Nanotechnol. 8, 199–205 (2013).
sensitive and prolonged intracellular measurement of 151. Lee, J. H. et al. Magnetic nanoparticles for ultrafast 178. Kim, D. H. et al. Biofunctionalized magnetic-vortex
action potentials. Nat. Commun. 5, 3206 (2014). mechanical control of inner ear hair cells. ACS Nano 8, microdiscs for targeted cancer-cell destruction. Nat.
123. Almquist, B. D. & Melosh, N. A. Fusion of biomimetic 6590–6598 (2014). Mater. 9, 165–171 (2010).
stealth probes into lipid bilayer cores. Proc. Natl Acad. 152. Pankhurst, Q. A., Thanh, N. T. K., Jones, S. K. & 179. Mannix, R. J. et al. Nanomagnetic actuation of
Sci. USA 107, 5815–5820 (2010). Dobson, J. Progress in applications of magnetic receptor-mediated signal transduction. Nat.
124. Patolsky, F. et al. Detection, stimulation, and inhibition nanoparticles in biomedicine. J. Phys. D: Appl. Phys. Nanotechnol. 3, 36–40 (2008).
of neuronal signals with high-density nanowire 42, 224001 (2009). 180. Carrey, J., Mehdaoui, B. & Respaud, M. Simple
transistor arrays. Science 313, 1100–1104 153. Guduru, R. et al. Magneto-electric nanoparticles to models for dynamic hysteresis loop calculations of
(2006). enable field-controlled high-specificity drug delivery to magnetic single-domain nanoparticles: application to
125. Tian, B. & Lieber, C. M. Synthetic nanoelectronic eradicate ovarian cancer cells. Sci. Rep. 3, 2953 magnetic hyperthermia optimization. J. Appl. Phys.
probes for biological cells and tissues. Annu. Rev. (2013). 109, 083921 (2011).
Anal. Chem. 6, 31–51 (2013). 154. Lugo, K. et al. Remote switching of cellular activity 181. Chen, R. et al. Wireless magnetothermal deep brain
126. Duan, X. et al. Intracellular recordings of action and cell signaling using light in conjunction with stimulation. Science 347, 1477–1480 (2015).
potentials by an extracellular nanoscale field-effect quantum dots. Biomed. Opt. Express 3, 447–454 182. Huang, H. et al. Remote control of ion channels and
transistor. Nat. Nanotechnol. 7, 174–179 (2012). (2012). neurons through magnetic-field heating of
127. Tian, B. et al. Three-dimensional, flexible nanoscale 155. Pappas, T. C. et al. Nanoscale engineering of a cellular nanoparticles. Nat. Nanotechnol. 5, 602–606
field-effect transistors as localized bioprobes. Science interface with semiconductor nanoparticle films for (2010).
329, 830–834 (2010). photoelectric stimulation of neurons. Nano Lett. 7, The first application of MNP heating to regulate
128. Liu, J. et al. Syringe-injectable electronics. Nat. 513–519 (2007). intracellular calcium with AMFs.
Nanotechnol. 10, 629–636 (2015). 156. Rowland, C. E. et al. Electric field modulation of 183. Stanley, S. A. et al. Radio-wave heating of iron oxide
129. Xie, C. et al. Three-dimensional macroporous semiconductor quantum dot photoluminescence: nanoparticles can regulate plasma glucose in mice.
nanoelectronic networks as minimally invasive brain insights into the design of robust voltage-sensitive Science 336, 604–608 (2012).
probes. Nat. Mater. 14, 1286–1292 (2015). cellular imaging probes. Nano Lett. 15, 6848–6854 184. Nair, M. et al. Externally controlled on demand release
130. Fu, T. M. et al. Stable long-term chronic brain mapping (2015). of anti-HIV drug using magneto-electric nanoparticles
at the single-neuron level. Nat. Methods 13, 157. Marshall, J. D. & Schnitzer, M. J. Optical strategies for as carriers. Nat. Commun. 4, 1707 (2013).
875–882 (2016). sensing neuronal voltage using quantum dots and 185. Guduru, R. et al. Magnetoelectric ’spin’ on stimulating
A pioneering demonstration of long-term other semiconductor nanocrystals. ACS Nano 7, the brain. Nanomedicine 10, 2051–2061 (2015).
recordings using microstructured electrode meshes 4601–4609 (2013). 186. Shin, T. H. et al. Recent advances in magnetic
with a negligible tissue response. 158. Derfus, A. M., Chan, W. C. & Bhatia, S. N. Probing the nanoparticle-based multi-modal imaging. Chem. Soc.
131. Wang, Y. & Guo, L. Nanomaterial-enabled neural cytotoxicity of semiconductor quantum dots. Nano Rev. 44, 4501–4516 (2015).
stimulation. Front. Neurosci. 10, 69 (2016). Lett. 4, 11–18 (2004). 187. Davies, G.-L., Kramberger, I. & Davis, J. J.
132. Gao, J., Gu, H. & Xu, B. Multifunctional magnetic 159. Murphy, C. J. et al. Gold nanoparticles in biology: Environmentally responsive MRI contrast agents.
nanoparticles: design, synthesis, and biomedical beyond toxicity to cellular imaging. Acc. Chem. Res. Chem. Commun. (Camb.) 49, 9704–9721 (2013).
applications. Acc. Chem. Res. 42, 1097–1107 (2009). 41, 1721–1730 (2008). 188. Atanasijevic, T. et al. Calcium-sensitive MRI contrast
133. Hu, M. et al. Gold nanostructures: engineering their 160. Yoo, S. et al. Photothermal inhibition of neural activity agents based on superparamagnetic iron oxide
plasmonic properties for biomedical applications. with near-infrared-sensitive nanotransducers. ACS nanoparticles and calmodulin. Proc. Natl Acad. Sci.
Chem. Soc. Rev. 35, 1084–1094 (2006). Nano 8, 8040–8049 (2014). USA 103, 14707–14712 (2006).

NATURE REVIEWS | MATERIALS VOLUME 2 | ARTICLE NUMBER 16093 | 15


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
REVIEWS

189. Shapiro, M. G. et al. Dynamic imaging with MRI 201. van Landeghem, F. K. H. et al. Post-mortem studies in administration. J. Spinal Cord Med. 30, 477–481
contrast agents: quantitative considerations. Magn. glioblastoma patients treated with thermotherapy (2007).
Reson. Imaging 24, 449–462 (2006). using magnetic nanoparticles. Biomaterials 30, 213. Kowal, S. L. et al. The current and projected
190. Rodriguez, E. et al. Magnetic nanosensors optimized 52–57 (2009). economic burden of Parkinson’s disease in the United
for rapid and reversible self-assembly. Chem. Commun. 202. Tang, J. C. et al. A nanobody-based system using States. Mov. Disord. 28, 311–318 (2013).
(Camb.) 50, 3595–3598 (2014). fluorescent proteins as scaffolds for cell-specific gene 214. Calabresi, P. et al. Direct and indirect pathways of
191. Lee, T., Cai, L. X., Lelyveld, V. S., Hai, A. & Jasanoff, A. manipulation. Cell 154, 928–939 (2013). basal ganglia: a critical reappraisal. Nat. Neurosci.
Molecular-level functional magnetic resonance 203. Antman-Passig, M. & Shefi, O. Remote magnetic 17, 1022–1030 (2014).
imaging of dopaminergic signaling. Science 344, orientation of 3D collagen hydrogels for directed 215. Arber, S. Motorcircuits in action: specification,
533–535 (2014). neuronal regeneration. Nano Lett. 16, 2567–2573 connectivity, and function. Neuron 74, 975–989
192. Christiansen, M. G. et al. Magnetically multiplexed (2016). (2012).
heating of single domain nanoparticles. Appl. Phys. 204. Pakulska, M. M., Ballios, B. G. & Shoichet, M. S. 216. Holstege, G. Descending motor pathways and the
Lett. 104, 213103 (2014). Injectable hydrogels for central nervous system spinal motor system: limbic and non-limbic
193. Wijaya, A. et al. Selective release of multiple DNA therapy. Biomed. Mater. 7, 024101 components. Prog. Brain Res. 87, 307–421
oligonucleotides from gold nanorods. ACS Nano 3, (2012). (1991).
80–86 (2008). 205. Shapiro, M. G. et al. Biogenic gas nanostructures as 217. Hammerle, H. et al. Biostability of micro-photodiode
194. Chen, R. et al. High-performance ferrite nanoparticles ultrasonic molecular reporters. Nat. Nanotechnol. 9, arrays for subretinal implantation. Biomaterials 23,
through nonaqueous redox phase tuning. Nano Lett. 311–316 (2014). 797–804 (2002).
16, 1345–1351 (2016). 206. Stanley, S. A. et al. Bidirectional electromagnetic 218. Patrick, E. et al. Corrosion of tungsten microelectrodes
195. Chen, Y. & Liu, L. Modern methods for delivery of control of the hypothalamus regulates feeding and used in neural recording applications. J. Neurosci.
drugs across the blood–brain barrier. Adv. Drug Deliv. metabolism. Nature 531, 647–650 Methods 198, 158–171 (2011).
Rev. 64, 640–665 (2012). (2016). 219. Wang, A. et al. Stability of and inflammatory response
196. Salvati, A. et al. Transferrin-functionalized 207. Wheeler, M. A. et al. Genetically targeted magnetic to silicon coated with a fluoroalkyl self-assembled
nanoparticles lose their targeting capabilities when a control of the nervous system. Nat. Neurosci. 19, monolayer in the central nervous system. J. Biomed.
biomolecule corona adsorbs on the surface. Nat. 756–761 (2016). Mater. Res. A. 81, 363–372 (2007).
Nanotechnol. 8, 137–143 (2013). 208. Chasteen, N. D. & Harrison, P. M. Mineralization in 220. Li, W. et al. Corrosion behavior of parylene–metal–
197. Tenzer, S. et al. Rapid formation of plasma protein ferritin: an efficient means of iron storage. J. Struct. parylene thin films in saline. ECS Trans 11, 1–6 (2008).
corona critically affects nanoparticle pathophysiology. Biol. 126, 182–194 (1999).
Nat. Nanotechnol. 8, 772–781 (2013). 209. Meister, M. Physical limits to magnetogenetics. eLife Acknowledgements
198. Kim, J. A. et al. Role of cell cycle on the cellular uptake 5, e17210 (2016). P.A. is supported by the National Science Foundation (NSF)
and dilution of nanoparticles in a cell population. Nat. 210. Komeili, A. et al. Magnetosomes are cell membrane through a CAREER Award, Center for Materials Science and
Nanotechnol. 7, 62–68 (2012). invaginations organized by the actin-like protein Engineering, Center for Sensorimotor Neural Engineering,
199. Hynynen, K. et al. Local and reversible blood–brain MamK. Science 311, 242–245 (2006). National Institutes for Neurological Disorders and Stroke,
barrier disruption by noninvasive focused ultrasound 211. Kolinko, I. et al. Biosynthesis of magnetic National Institute of Mental Health, the Defense Advanced
at frequencies suitable for trans-skull sonications. nanostructures in a foreign organism by transfer of Research Projects Agency, Dresselhaus Fund Award, and the
Neuroimage 24, 12–20 (2005). bacterial magnetosome gene clusters. Nat. Bose Research Grant.
200. Chertok, B. et al. Iron oxide nanoparticles as a drug Nanotechnol. 9, 193–197 (2014).
delivery vehicle for MRI monitored magnetic targeting 212. French, D. D. et al. Health care costs for patients with Competing interests statement
of brain tumors. Biomaterials 29, 487–496 (2008). chronic spinal cord injury in the veterans health The authors declare no competing interests.

16 | ARTICLE NUMBER 16093 | VOLUME 2 www.nature.com/natrevmats


©
2
0
1
7
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like