You are on page 1of 27

REVIEW

www.afm-journal.de

Innovations in 3D Tissue Models of Human Brain


Physiology and Diseases
Michael L. Lovett, Thomas J. F. Nieland,* Yu-Ting L. Dingle, and David L. Kaplan*

specific cell-types, and they may be related


3D laboratory tissue cultures have emerged as an alternative to traditional to the organization of cis-regulatory ele-
2D culture systems that do not recapitulate native cell behavior. The discrep- ments in DNA or stress conditions such
ancy between in vivo and in vitro tissue-cell-molecular responses impedes as aging.[8–15] Ultimately, they may lead to
understanding of human physiology in general and creates roadblocks for the changes in interactome at the mesoscale,
such as lipid profiles or the composition
discovery of therapeutic solutions. Two parallel approaches have emerged for
of the synaptic proteome.[16,17] In light of
the design of 3D culture systems. The first is biomedical engineering meth- this broad spectrum of species-specific dif-
odology, including bioengineered materials, bioprinting, microfluidics, and ferences on the level of DNA, molecular
bioreactors, used alone or in combination, to mimic the microenvironments pathways, cells, and brain networks, it
of native tissues. The second approach is organoid technology, in which stem should not come as a surprise that the
cells are exposed to chemical and/or biological cues to activate differentiation modeling in rodents of complex neurode-
generative diseases and psychiatric disor-
programs that are reminiscent of human (prenatal) development. This review ders pose challenges, in particular for drug
article describes recent technological advances in engineering 3D cultures and target discovery.[3,5,18] Unlike rodent
that more closely resemble the human brain. The contributions of in vitro 3D models, access to living human test sub-
tissue culture systems to new insights in neurophysiology, neurological dis- jects and primates, or ex vivo brain slices
eases, and regenerative medicine are highlighted. Perspectives on designing derived thereof, is surrounded with ethical
issues. Moreover, only a limited number
improved tissue models of the human brain are offered, focusing on an inte-
of scientists and clinicians master the nec-
grative approach merging biomedical engineering tools with organoid biology. essary expertise in, or have access to, brain
imaging techniques such as noninvasive
functional magnetic resonance imaging
1. Introduction and electroencephalography, or minimally invasive methods
such as the implantation of recording and stimulating elec-
To understand human brain physiology, the human condition trodes in the brain. These methods also suffer from low spatial
must be studied in a laboratory setting. Animal models, such and/or temporal resolution, as well as the need for large patient
as rodents, primates, worms and experimental systems such as cohorts and consortia to identify the features of the human
yeast, have yielded great insights into development, function, brain connectome related to normal aging or disease.[19–22]
and diseases of the brain. However, these model systems do not The advent of stem cell engineering, both embryonic (ESC)
fully recapitulate the complexity of the human brain related to and induced pluripotent stem cells (iPSC), has been a game
neural circuit architecture, genetics, gene expression, cell-type changer in the study of the human brain. These advancements
diversification, lipidomics, and proteomics. The cerebral cortex, in cell biology made possible, for the first time, the generation of
the part of the brain involved in motor and sensory function a renewable source (vis-à-vis terminal explant studies) and propa-
and processing information, as well as associated with psy- gation of in vitro human culture models of the human brain that
chiatric diseases, is less complex in rodents in comparison to are scalable and available to the scientific community at large.[23]
primates in general, and to humans in particular.[1–5] Evolution iPSCs, which are generated from healthy individuals or patient
also had its impact on the divergence of genetic make-up of fibroblasts or their peripheral blood monocytes, have demon-
humans compared to other mammals, notably the expression strated to be excellent tools to study human neurological diseases.
of noncoding RNAs.[6,7] Species-specific gene expression pat- The literature is now replete with examples from Alzheimer’s
terns are found in the context of neural circuit function and and Parkinson’s disease, schizophrenia, autism, epilepsy, pain,
and other fields.[23] Genome engineering approaches, in par-
Dr. M. L. Lovett, Prof. T. J. F. Nieland, Dr. Y.-T. L. Dingle, Prof. D. L. Kaplan ticular based on clustered regularly interspaced short palin-
Department of Biomedical Engineering dromic repeats technology (other examples are zinc-fingers
Tufts University and transcription activator-like effector nucleases), have
4 Colby Street, Medford, MA 02155, USA made a similar transformative impact. Genetic variants associ-
E-mail: thomas.nieland@tufts.edu; david.kaplan@tufts.edu
ated with increased (or decreased) risk of disease can now be
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adfm.201909146.
introduced, with relative ease, in the germ-line of healthy cells
to replicate human disease. They can also be removed from the
DOI: 10.1002/adfm.201909146 cells as a strategy to correct brain illnesses.[24–28] These cells may

Adv. Funct. Mater. 2020, 1909146 1909146  (1 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

be used in traditional in vitro mono-cultures grown on flat, 2D


surfaces, but these experimental cell systems do not replicate the Thomas Nieland is a
complexity and biology of the human brain. Salient examples of Research Associate Professor
this fact can be found in studies of gene expression signatures, in the Department of
extracellular matrix (ECM) microenvironment, neural network Biomedical Engineering at
firing, and biomarkers of brain development and disease.[29–34] Tufts University. His research
This realization has spurred the development of 3D laboratory interest is to understand
cell and tissue models that aim to better replicate the cellular the interactions between
composition, microenvironment, and architecture of the brain. different brain cell-types
We will discuss two different techniques that have become and neural circuits that may
popular for designing 3D brain-like tissue structures. The first is lead to neurodegenerative
the use of organoids (grown in bioreactors or otherwise), dense and psychiatric diseases.
multicellular tissues differentiated from stem cells following Bringing unique aca-
their innate developmental cues.[35–37] The other is bioengi- demic and industry experiences, his ultimate goal is drug
neering, a term collectively used to describe methods that make discovery using a combination of bioengineering with stem
use of a bottom up strategy, involving seeding (multiple) cells cell technology and systems biology.
(types) in biomaterial scaffolds and/or hydrogel environments,
including those generated by 3D printing, and the use of bio-
reactors such as microfluidic devices or perfusion devices.[38,39] David Kaplan is the Stern
These two complementary approaches have matured to the Family Endowed Professor of
stage that we now see bioengineering techniques being used in Engineering at Tufts University
a complementary fashion to improve on organoid cultures.[40] and a distinguished university
We will conclude with future perspectives on generating more professor. He is Professor
complex in vitro tissue models of the human brain that inte- and Chair of the Department
grate organoid and bioengineering technology, with a vision of Biomedical Engineering.
toward high-throughput screening systems for biomarker and His research focus is on
therapeutic discovery. biopolymer engineering to
understand structure–function
relationships for biomaterials,
tissue engineering, and regen-
2. Ex Vivo and 2D In Vitro Platforms erative medicine.
to Study the Brain
Much of our fundamental understanding of the brain is derived
from traditional in vitro culture systems. Through extensive technological capabilities, postmortem brain studies continue
research studies of postmortem brain tissues as well as isolated to be an indispensable tool to study human brain function
cell cultures in 2D, many key features of brain physiology and and disorders through the ability to observe and analyze tissue
mechanisms of disease have been identified. This knowledge structures, cellular morphologies, and protein and gene expres-
serves as the basis for generating 3D models of the brain, and sion.[42] As an example, the examination of AD patient brain
thus we begin by reviewing the literature on traditional in vitro tissues has been used to identify what are now described as the
systems to provide context to the sections describing 3D brain two classical molecular hallmarks of the disease—extracellular
models. accumulation of beta-amyloid (Aβ) plaques and intracellular
aggregation of hyperphosphorylated tau proteins in neurofi-
brillary tangles.[43] More recently, transcriptomics analysis and
2.1. Human Postmortem Brain Tissues single cell-RNA sequencing from AD brain tissues have identi-
fied region-specific and cell-type-specific expression patterns.[43]
Historically, preliminary attempts to understand the struc- By taking advantage of new technologies such as single-cell
ture and function of the brain, the most complex organ in RNA sequencing, a reduction in excitatory neurons in the AD-
the body, have relied on the use of postmortem tissues. In a gene carrier but not in unaffected family members or sporadic
classic example from the 1860s, the French physician Pierre cases has been found, indicating differences in disease mecha-
Paul Broca found lesions in the same area of the postmortem nisms between familiar and sporadic forms of the disease.[44]
brains of 12 patients with impaired language skills, providing Similarly, combinations of postmortem studies detailed the
the first evidence linking particular brain regions to specific molecular changes in the PD midbrain, including abnormal
functions.[41] Since then, many other key discoveries have been levels of α-synuclein production and aggregation, reduction in
made by examining and identifying the hallmark anatomical dopamine synthesis, increased oxidative stress, mitochondrial
changes in postmortem brains associated with patients who dysfunction, and reduction in neurotrophic factors essential
suffer from neurological dysfunctions such as Alzheimer’s for cell survival.[45] A recent study closely examined a small
disease (AD; loss of neurons in cortex and subcortical regions) number of postmortem brain tissues of patients who received
and Parkinson’s disease (PD; loss of dopamine neurons in the deep brain stimulation therapy, an approved surgical interven-
substantia nigra of the midbrain). Due to advances in tion alternative to medication. The results suggested recovery of

Adv. Funct. Mater. 2020, 1909146 1909146  (2 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

mitochondria numbers and volume as underlying mechanisms native cell types, ECM, cell organization, tissue architecture,
of neuroprotection.[46] and local synaptic connections. In addition, multiple tissue
In a recent and high-profile example of identifying neurode- samples are generated from each brain, reducing the number
generative disease through examination of postmortem tissues, of animals needed. In typical organotypic slice studies, brain
the study of brains subjected to repeated injuries, particularly slices are prepared with a thickness of 100–400 µm from
from American football players, has led to the identification of acutely sacrificed animals.[55] These slices can be used in acute
chronic traumatic encephalopathy (CTE). Using postmortem experiments, in which they are analyzed immediately after
brain tissues, researchers were able to define the distinguish- dissection to provide information from a near-in vivo state, or
able molecular feature of perivascular tau protein pathology. cultured in media for longer term studies (days to months) for
A postmortem study on four teenage athletes’ brains during the purpose of studying cell behavior over time after in vitro
the acute and subacute (1 day to 4 months) phase post brain stimuli. A recent study compared the level of membrane associ-
trauma showed microvascular damage, neuroinflammation, ated tau protein versus tau protein released into the media in
axonal injuries, and tau pathology—indication of early stage of brain slice cultures from wild type and AD model mice over
CTE.[47] the duration of 1 month.[56] A higher level of tau protein was
Postmortem brains have similarly been used to study released from the AD slices and, unlike in wild type slices, the
neurodevelopmental and neuropsychiatric disorders, such as release did not respond to neuronal stimulation. These results
autism spectrum disorder (ASD), schizophrenia, and bipolar suggested differences in the mechanisms in tau protein release
disorder. These disorders are particularly challenging as the and spreading in AD.[56]
majority of cases have no known genetic cause. To begin to Organotypic slices are particularly useful when the tissue
address this issue, hundreds of postmortem brains from ASD structure and organization are critical to the study outcome,
patients and control cohorts were recently investigated, with such as in traumatic brain injury (TBI), cancer biology, or
allelic imbalance (expression favors one allele) reported in two neural circuit structure and functional activity. The tearing
specific regions associated with small nucleolar RNA.[48] More and shearing of the soft, connected brain tissue as a conse-
recent studies examining large numbers of tissue samples and quence of TBI, is not sufficiently recapitulated by single cells
data sets have identified the association of neuroinflamma- cultured on a conventional 2D hard plastic surface, but can be
tion with mental illnesses, such as schizophrenia and bipolar mimicked by mechanically stretching the brain slices. Comple-
disorder.[49–51] mentary to in vivo mouse models, slice culture allows for the
As evidenced above, the precise genetics, cell types tracking of cell death, tissue injury, inflammation, and changes
involved, and pathogenesis are not well understood for most in neural network activity over time.[57,58] Brain slice models of
brain disorders. Phenotype heterogeneity in presentations TBI also provide a useful tool for screening therapeutic strate-
and confounding factors such as sex, age, medications, and gies, and several candidates have already been identified, such
environmental factors add to the difficulties in investigating as hypothermia,[59] inert gases,[60] sodium calcium exchanger
brain disorders.[52] Also limiting is the availability of post- blockers,[61,62] and glutamate receptor blockers.[63] Although
mortem brain samples. Recently, successful efforts to increase hypothermia has been used in the treatment for TBI, a recent
awareness for brain donations has made possible the building review did not identify improvement in clinical outcomes.[64]
of networks of brain tissue repositories and databases. The Because tumor metastasis to the brain and brain tumor inva-
NIH NeuroBioBank (NBB) networks six brain tissue reposi- siveness directly correlates to poor prognosis, many researchers
tories, with the largest one storing over 2000 brain samples, study tumor cell migration in vitro. The majority of these
in order to facilitate the collection and redistribution of high- studies, however, utilize 2D scratch or transwell assays that
quality brain tissues in the United States.[53] capture few of the actual in vivo interactions between tumor
Despite the high level of relevance to human brain disorders, cells, host cells and ECM, or the cellular heterogeneity of the
postmortem brain studies are limited in their ability to provide cancer. Moreover, the majority of organotypic slice studies use
information on dynamic, temporal changes of disease progres- rodent brains, and therefore they face the same drawbacks of
sion. Furthermore, from the perspective of disease prevention lacking proper human-equivalent models as other brain dis-
or treatment, molecular and cellular changes may occur years orders. There are slice culture models that attempt to address
prior to clinical presentation of symptoms of the decline in these critical issues by employing the implantation of human
cognitive or motor skills[54] and researchers cannot control the tumor cell lines in the form of single cells or spheroids on
genetic background or the external insults to the study objects. rodent brain slices. This made it possible to examine in detail
Thus, the need for well-controlled laboratory models are essen- the interactions between glioblastoma cells and the vascula-
tial for probing the human brain (Figure 1). ture, identifying surface protein targets and small molecule
candidates to reduce tumor invasion or migration. Further, this
approach provided insight into changes in gene expression in
2.2. Organotypic Slice Culture from Animals reactive astrocytes, turning the microenvironment to become
pro-metastasis.[65–67]
In contrast to human post mortem tissues, in vitro organotypic In any slice culture, one of the key limitations is the age-
slice cultures from animals allow for the control of genetics, dependency of slice survival and the resulting impact on study
sex, age, external insults, and reduction of confounding factors. design. Slices have high viability when prepared from embry-
The major advantage of in vitro organotypic slice cultures is that onic animal brains, but at the same time contain immature
they provide investigators access to the cells while preserving cell and phenotype. Slices from postnatal animals provide

Adv. Funct. Mater. 2020, 1909146 1909146  (3 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 1.  Overview of commonly used model systems for studying the brain.

a higher level of cellular and tissue maturation than from fibronectin, Matrigel, poly-lysine, and/or poly-L-ornithine for
embryonic animals and are more resistant to the mechanical adhesion.[73] In comparison to the abovementioned models, cell
trauma from slice preparation (up to months).[68] Adult brain culture allows for controlled co-cultures with different cell types
slices, including from animals and from human biopsy and and/or cells with different genetic backgrounds (Table 1).
postmortem tissues, historically have a poor survival rate, but To date, investigating the human brain in vitro has been hin-
improvements have been made by adding neurotrophic factors, dered by the inaccessibility to live human brain tissue and the
the addition of cerebrospinal fluid in the medium, or media extreme difficulty of isolating and culturing primary human
flow systems to enhance nutrient and oxygen exchange.[69–72] neurons. Historically, rodent embryonic or postnatal neuronal
However, it should be noted that slice preparation induces sig- culture has been used for studying neuronal functions and to
nificant mechanical trauma to the tissue, which can cause the model diseases. For these studies, the neurons are often geneti-
tissue to have injury-like responses that may affect the outcome cally modified or cultured in the presence of toxins or disease-
of the experiment and introduce artifacts to interpretations into associated protein, such as Aβ for AD. Given that the underlying
brain functions. etiologies are not well understood for many human neurodevel-
opmental, neuropsychiatric, and neurodegenerative disorders,
there is a lack of appropriate rodent models and the subsequent
2.3. 2D Cell Culture Systems results from rodent cell culture often fail to translate to the
human condition. A major breakthrough toward this goal was
Culturing dissociated cells in vitro offers the advantages of the method of reprogramming somatic cells to iPSCs, developed
optical transparency, substrate consistency, easy handling, by Takahashi and Yamanaka in 2006, achieved by overex-
robustness, reproducibility, relatively low cost, compat- pressing transcription factors Oct3/4, Sox2, Klf4, and c-Myc in
ibility with electrophysiology techniques, and high-throughput fibroblasts. Since then, many protocols have been developed for
screening. In conventional 2D culture, dissociated cells are the direct differentiation of iPSCs or the transdifferentiation
cultured on plastic or glass surfaces coated with laminin, into relevant brain cell types, including cortical neurons, striatal

Adv. Funct. Mater. 2020, 1909146 1909146  (4 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Table 1.  Summary of strengths and limitations of ex vivo and 2D in vitro platforms for brain studies.

Strengths Limitations
[41–54]
Human postmortem tissues Complexity of brain preserved, including cell composition Tissue not living, no evaluation of function or development
and spatial organization over time
Identification of disease markers Limited reproducibility or scalability for high-throughput screening
Imaging of brain structure
Rodent slice cultures[55–72] Complexity of brain preserved, including cell composition Limited capability for evaluating function or
and spatial organization development over time
Analysis of neural network anatomy and function Limited reproducibility or scalability
Not suitable for high-throughput screening
2D cell culture[73–86] Reproducible and scalable: high-throughput screening Limited culture longevity, particularly for primary cells as opposed
to stem cells
Suitable for noninvasive imaging and electrophysiological recordings Low complexity, not representative of native cell environment
Modeling of human biology through stem cell technology

dopamine neurons, astrocytes, microglia, oligodendrocytes, an expansion of CAG repeats in the huntingtin (HTT) gene,
pericytes, and brain microvasculature endothelial cells.[74,75] iPSC models have been used to identify the genetic and cellular
The use of iPSCs from individuals affected by neurological mechanisms of disease progression. A recent study of patient
disorders presents an unprecedented tool to unveil the under- iPSC-derived cortical neurons found modifications in transcrip-
lying genetics, mechanisms, and pathogenesis associated with tomics, morphology, and maturation, suggesting altered corti-
these disorders. Within the past decade, iPSC-derived neurons cogenesis.[84] Another study detected astrocyte dysfunction in a
have provided novel insights on the etiology and demonstrated HD-iPSC co-culture model, suggesting contribution of glial–
opportunities to screen for therapeutic interventions. neuron interactions in neuronal death and the importance of
Due in part to the fact that AD is the most prevalent neu- studying multiple cell types in a disease model.
rodegenerative disorder, many iPSC lines have been derived Since the timeline for iPSC differentiation and maturation
from AD patients, and their healthy family members, for the into neurons or other cell-types follows the same trajectory as
study of monogenetic and sporadic forms of the disease, as in the developing embryo, these cells are particularly useful
well as novel candidate genetic risk factors identified in genetic tools to study neurodevelopmental disorders such as ASDs.
analysis of patients.[75–78] Studies have shown that human Several iPSC lines have been established from monogenic
iPSC (hiPSC)-derived cortical neurons carrying familial ASD-iPSC (e.g., Fragile X syndrome, Rett syndrome, Timothy
Alzheimer’s disease (fAD) mutations presented elevated levels syndrome, Angelman syndrome, and Phelan-McDermid syn-
of Aβ42—the main component of Aβ plaque, as well as patho- drome) as well as from idiopathic ASD patients. Investiga-
logical tau protein phenotype and neuroinflammation as seen tors use these ASD-iPSC-neurons to study cell proliferation,
in patients, demonstrating the utility of iPSCs as a preclinical excitatory and inhibitory balance, synapse connectivity, neural
model of the disease.[79–81] Treatment with pharmacological networks, epigenetics, and proteomics.[85] Many of them
compounds targeting pathways related to Aβ production and showed upregulation of inhibitory neuron genes, deficits in
cleavage reduced AD pathology, supporting the feasibility of excitatory synaptic transmission, reduction in neurite length
iPSC models as in vitro platforms for identifying therapeutic and synaptogenesis, indicating the relevance of abnormal net-
interventions. work formation in ASD.[85]
These approaches have been used to study the aspects of The development of cells and 2D cell culture models has led
other neurodegenerative disorders such as Parkinson’s or Hun- to tremendous discoveries in the field. Despite this insight, the
tington diseases. One of the hallmarks of Parkinson’s disease main drawback of 2D cell culture remains that they are oversim-
is death of midbrain dopamine neurons and the formation of plified and lack the complexity and organization of the in vivo
α-synuclein protein aggregates. Several different types of PD microenvironment (Figure  2). As an example, the architecture
models have been established, either by treating healthy iPSC- of the original neural networks is lost when neurons are disso-
derived dopamine neurons with α-synuclein aggregates or by ciated into single cells. Furthermore, cell density in 2D culture
generating DA neurons from familial PD (fPD) and sporadic is typically much lower than that found in vivo, with cell–cell
PD (sPD) patients. iPSC lines derived from rare fPD cases interactions limited to the XY directions. The local cellular envi-
have helped in the validation of novel risk genes.[82] One study ronment is also altered in 2D as typical cell culture substrates
compared neurons derived from sPD and fPD iPSCs, inter- are several orders of magnitude more rigid than brain tissues,
estingly finding epigenetic dysregulation in both groups with and the dynamics of nutrient, oxygen, and secreted factor trans-
DNA methylation patterns differing from healthy dopamine port are much different in open media than in 3D tissues. As
neurons and hypermethylation in regions relevant to PD.[83] an example, an AD model with iPSC-neurons cultured in 2D
For the study of Huntington’s disease (HD), a hereditary, were unable to generate Aβ plaques as efficiently as in 3D
monogenic, neurodegenerative movement disorder caused by because the secreted Aβ freely diffused away.[86] Because many

Adv. Funct. Mater. 2020, 1909146 1909146  (5 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Subsequent large-scale RNA sequence analysis of thousands


of single cells have expanded on the notion that organoids give
rise to a broad range of cell-types, such as neuroepithelial cells,
photoreceptors, and dopamine neurons.[94] Gene expression,
epigenetic, and image analysis studies further indicate that the
maturing organoids follow the normal pace of development of
the human fetal brain,[95–98] adhering to species-specific rules
of evolution.[99–101] As such, organoids have become exciting
tools to examine the evolutionary cues that shape development
of the brain, in particular for studies inspired by genetic anal-
ysis comparing humans to Neanderthals, chimpanzee, or other
animals.[102–105]

3.2. Optimizing Differentiation Protocols for Organoid Cultures

Organoid technology is often associated with limitations related


to reproducibility and heterogeneity between organoids from
the same or different batches as well as necrosis. As a conse-
Figure 2.  Overview of the applicability of in vitro platforms to emulate quence, the implementation of organoid cultures to study brain
brain physiology at different levels of brain organization. Bars represent development and function is challenging. To overcome these
the optimal application of each technology.
hurdles, several groups have developed the next generation of
protocols for culturing organoids that are more homogenous.
environmental factors can significantly impact cell behavior Commonly referred to as “directed differentiation,” these
and thus disease mechanisms, the development of 3D tissue methods make use of cocktails of chemicals and trophic factors,
model systems is actively pursued. either alone[106] or in combination with low-adhesion plates,
and custom-built miniaturized bioreactors or 3D scaffold envi-
ronments.[107] With these techniques, it is now possible to guide
3. 3D Brain Organoid Models the development of iPSCs toward specific sublineages, such as
forebrain, midbrain, or hypothalamus.[98,108–117] The success of
3.1. Organoids as an Emerging Technology for Analyzing the these efforts in achieving more reliable organoid differentiation
Human Brain was demonstrated recently in the testing of multiple iPSC lines,
independent differentiation, and analyzing hundreds of thou-
All methods for generating organoids rely on the innate sands of single cells and in 96-well plate format assays.[115,118,119]
properties of stem cells to develop and self-organize into 3D
architectures that share similarities with native tissues. Stem
cells are uniquely endowed with the capacity to self-renew, dif- 3.3. Organoids as a Tool to Study Neurodevelopment,
ferentiate, and mature into virtually any type of cell.[23] As such, Disease, and Evolution
ESCs and iPSCs are ideal substrates for bioengineering of orga-
noids. The first organoid systems generated modeled peripheral The power of organoid technology for characterization of
(i.e., not the central nervous system) tissues using mouse, and neurological disease signatures was first demonstrated in a
then human, cells. Salient examples derive from the retina[87,88] patient population stratified based on clinical record but with
and the anterior pituitary,[89] two tissue types composed of no knowledge of the genetic underpinnings. Focusing on idi-
neural and nonneural cell types, and the intestine.[90–92] opathic forms of ASD, differences in cell proliferation, neural
The first protocols for engineering brain organoids made use differentiation, and synapse biology were observed in organoids
of embryoid bodies (EBs) generated from hiPCSs and grown on of iPSCs selected from patients with macrocephaly phenotypes
mouse embryonic fibroblasts. These cells were then embedded compared to healthy controls. With “omics” analyses, changes
in Matrigel droplets and cultured for prolonged time (up to in a forkhead box G1 (FOXG1) transcription factor-dependent
10 months) in a spinning bioreactor. The resulting cerebral signaling was found associated with an imbalance between
organoid resembled the characteristics of the human cortex, gamma-aminobutyric acid (GABA)-ergic and glutamatergic
including the multilayered structure and inner and outer sub- neurons,[108] often cited as one of the underlying culprits of
ventricular zones, which house progenitor cells and (outer) ASD.[120] Interestingly, aberrant FOXG1 signaling is also impli-
radial glia. This type of glia plays an important role in human cated in Rett syndrome, a rare neurodevelopmental disorder
evolution, driving the expansion of the human brain compared mainly affecting girls.[121,123] Organoid studies link microRNA
to other primates. Generated from hiPSCs, the emerging 3D pathways to altered extracellular signal-regulated kinase (ERK)
systems were heterogenous, also showing characteristics of and serine/threonine-protein kinase (AKT) signaling pathways
meninges, choroid plexus, and retina structures.[35,93] A related involved in neurogenesis in patients with Rett syndrome.[124]
approach relied on low-adhesion plates instead of bioreactors, A unique approach to the study of human (brain) devel-
and embryonic stem cells rather than iPSCs.[37] opment is the comparison of our genome to closely related

Adv. Funct. Mater. 2020, 1909146 1909146  (6 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

species, such as Neanderthals and the chimpanzee. Such com- neuron content, an increase in the pathogenic alpha-synuclein
parative analysis led to an intriguing organoid study of NOTCH protein, and changes in gene expression that implicate the
signaling and regulators of the PI3K-AKT-mTOR pathway in lysosome.[154] In an interesting contrast with conventional 2D
radial glia, implicating the rapid evolution of the larger human cultures, only mid-brain style organoids produce neuromelanin
cortex to loss of genome stability and ensuing neurodevelop- in vitro, a biomarker for aging and PD that is produced by the
mental disorders, such as 1q21.1 deletion and duplication substantia nigra, the part of the brain affected early in PD.[31]
syndrome.[103,125] Related studies point to the importance of
specific sets of chromatin enzymes, gene networks regulated
by long noncoding RNAs and enhancers involved with brain 3.5. The Impact of Environmental Insults and Infection
development,[126] with a potential link to autism.[105,126–128] on Human Biology in Brain Organoids
Organoids have been used to successfully study a wide range
of neurological disease such as hereditary spastic paraplegia, Brain organoids are also useful tools for examining brain function
periventricular heterotopia, a disease associated with intellectual and related pathology that does not depend on a genetic compo-
disability and dysfunction of cadherin adhesion molecules,[129] nent per se. Organoids have been used to model changes in neu-
and autoimmune diseases caused by neuro-inflammation.[130] rodevelopment caused by hypoxia (and rescue thereof)[155,156] and
Organoid studies of neurodevelopmental diseases successfully exposure to trace elements,[157] anticancer agents,[158] substances
highlighted defects in a broad range of cell–molecular func- of abuse,[159–161] therapeutic drugs,[162] and natural products that
tions, such as oxidative stress,[131] vesicular traffic,[132] calcium purportedly improve brain function.[163] Brain organoids have
channels,[114] the expansion of DNA nucleotide repeats in the similarly proven to be a useful tool in elucidating the pathology
context of Huntington’s disease,[133] chromatin remodeling underlying prion[164] and viral infection, such as Herpes Sim-
enzymes,[134,135] mRNA chemical modifications,[136] and lipid plex,[165] Dengue,[166] and, in particular, Zika virus (see below).
homeostasis.[137] A particularly intriguing application, that is Studies indicate that Japanese encephalitis virus appears to
otherwise hard to conduct in actual living humans, it the real- have a preference for astrocytes and neuroprogenitor cells, in
time tracking of dynamic phenotypes in organoids, such as particular outer radial glial cells, inducing cell death, and
neuronal migration and cytoskeleton dynamics.[35,97,114,138–143] immune responses when telencephalic organoids mature.[167]
In the face of the global public health crises that ensued
after the re-emergence of the virus in the Western world in
3.4. Organoids as a Tool to Study Neurodegenerative Diseases 2015, organoids quickly became a popular model to study Zika
virus (ZIKV) related pathology. Organoid models were instru-
Despite their typical fetal or early post-natal state,[95,99,144,145] it mental in finding determinants of viral tropism and identi-
has been possible to successfully model phenotypes in orga- fying how the virus may negatively impact brain development
noids that are typically associated with older age human brains. by impeding neurogenesis and cortical layer development
Alterations in cortical structure, neurogenesis, cell prolifera- processes.[101,109,168–172] An interesting connection between
tion, and cell signaling related to schizophrenia, which emerges ZIKV infection and the genesis of other brain pathologies such
during adolescence, have been replicated in ESC and iPSC- as schizophrenia and intellectual disability was established,
type organoids.[131,146,147] Organoids from patients suffering pointing to changes in the DNA methylome of ZIKV-infected
from early-onset (EO) fAD or Down Syndrome (DS), and the neuroprogenitors, neurons, and astrocytes with associated alter-
feeding of pathogenic Aβ peptides, recapitulate the progressive ations in gene expression.[173] Organoids have also been used in
neuropathology of patients, with the appearance first of Aβ and the search for therapeutic solutions for Zika virus, leading to
subsequently tau pathology[148–150] and inflammatory and ECM the discovery of key players in cellular entry of ZIKV[174,175] and
responses.[151] Intriguingly, DS-derived organoids also show small molecules with potential clinical benefits.[175–179] ZIKV
an increase in subclasses of GABAergic interneurons through triggers the endogenous RNAi-based anti-viral response, and
overexpression of OLIG2, providing one potential explanation broad-spectrum antibiotics such as enoxacin, which stimulates
for impaired cognitive ability of patients.[152] RNAi, abolishes viral infections and the ensuing microcephalic
Further mirroring clinical observations, the development of morphology of cerebral organoids.[180]
AD molecular pathology was shown to be delayed in organoids
expressing the late-onset (LO) E4 risk variant of a gene called
apolipoprotein E (APOE) relative to organoids from early-onset 3.6. Modeling Interactions between Different Cell-Subtypes
patients.[153] Related studies have highlighted a disease pathway and Different Brain Regions
shared between EO-fAD and LO-fAD involving gene networks
regulated by the repressor element-1 silencing transcription The circuitry of the brain is complex. Connections form
factor (REST).[154] In one of the first forays in using organoids between local and distant sub-networks, involving the commu-
to profile responses to therapeutic drugs, treatment with beta nication between many neuronal subtypes and support cells.[1,2]
and gamma-secretase inhibitors reduced amyloid and tau The functional maturation of developing organoids is now well
phenotypes associated with AD.[148] established through computational, fluorescence imaging, and
In PD, midbrain organoids from patients expressing the electrophysiology techniques.[181,182] Recent data suggest that
G2019S mutation in the leucine-rich repeat kinase 2 (LRRK2), the activity patterns emerging in organoids are more irregular
one of the prominent genetic causes of the disease, faithfully in time and space than conventional 2D-cultures, in a way more
reproduced hallmarks of the disease, such as reduced dopamine akin to that of the human brain.[33]

Adv. Funct. Mater. 2020, 1909146 1909146  (7 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

It is possible to generate functional organoids and use them 4. Engineering Approaches to 3D Brain Models
as a source to generate, isolate, and study a broad range of
cell subtypes, and to model the interactions between different In order to address the limitations of the systems described ear-
types of cells, including cancer cells,[183–185] tissue elements lier, microfluidics and bioreactor systems are increasingly being
such as the blood–brain barrier (BBB),[186] or even disparate used as enabling technologies for precise manipulation of the
brain regions. Protocols for glutamatergic, midbrain, and cellular environment (Table  2). These microscale systems,
dopaminergic neurons, inhibitory neurons, astrocytes and now commonly referred to as “tissue-on-a-chip,” are used to
oligodendrocytes, thalamic neuron, and organoid models recapitulate the physiological function of tissues in vitro, with
have been published.[31,98,108,110,114,143,145,175,187–190] Long-term noteworthy success in creating tissue models as diverse as the
(20 month) cultivation of organoids created an environ- human lung,[196,197] intestine,[198] and liver,[199] among others.
ment supportive of the transition of astrocytes from fetal to Recently, these systems have incorporated multiple tissue
a more mature stage.[145] Microglia-like cells, the brain resi- types into microphysiological systems (MPS), or “body-on-a-
dent immune cells that are derived from the mesoderm lin- chip” devices, comprised of individual chambers that mimic
eage, were successfully grafted onto organoids.[191] Recent the physiological function of different tissues all connected to
literature suggests that it is possible that functional micro- mimic inter-tissue communication.[200,201] Here, we focus on
glia arise naturally in brain organoids, which are ectoderm brain-specific microfluidic systems, but a thorough examina-
derived, possibly from a small pool of mesodermal cells.[192] tion of all advances in tissue- and body-on-a-chip systems can
When derived from Alzheimer’s patients carrying defined be found expertly reviewed elsewhere.[202,203]
early-onset mutations, microglia exacerbated AD phenotypes
when grafted onto organoid cultures.[153] Furthermore, within
a single organoid, developing cells may specialize into regions 4.1. Microfluidic Culture Devices
resembling the make-up of specific subregions of the brain.
Organoids can be fused to create 3D structures that emu- The production of a tissue-on-a-chip system requires control
late connections between ventral-rostral and dorsal areas over pattern and flow in a microscale environment.[202] Bor-
of the cortex and with the medial ganglionic eminence and rowing concepts from the photolithographic etching methods
thalamus.[110,111,114,143,187,193] In one of the first steps toward used to make computer chips, the innovative use of elastomeric
regenerative medicine applications, human organoids grafted materials such as poly(dimethylsiloxane) (PDMS) allowed for
in mouse brains showed enhanced functional maturation and the design and fabrication of devices with micro- to nanoscale
integration with the host tissue compared to the implantation control of feature shapes and sizes.[204] Networks of microfluidic
of neuroprogenitor cells.[194,195] channels and/or wells designed using computer-assisted design

Table 2.  Summary of strengths and limitations of 3D in vitro platforms for brain studies.

Strengths Limitations
[35–37,87–195]
Organoids Complexity of brain achieved through cell-guided tissue Limited control of cell composition and spatial organization
organization
Models of neurodevelopment, neurodegenerative disease, injury, Necrotic core limits tissue size
infection, evolution, etc.
Long-term culture capability and establishment of functional Low reproducibility and capability for high-throughput screening
networks
Microfluidic chips[196–253] Control of spatial organization and delivery of nutrients/growth Not suitable for long-term cultures to evaluate function or
factors via flow development
Capability for multi-tissue systems Limited capability for high-throughput screening applications
Integration with imaging or electrophysiological modalities Binding of drugs and biomolecules to polymer surfaces
Hydrogels[283–320,344–345,348–350,359–361] Ease of preparation for reproducible and scalable cultures Limited culture longevity due to hydrogel degradation
Suitable for noninvasive imaging and electrophysiological Variable ECM composition
recordings
Some control of spatial organization and size of tissues
Bioprinting[289,321–329,353–356] Control of spatial organization and size of tissues Requires optimization of cells and solution printing parameters
Suitable for noninvasive imaging and electrophysiological Limited culture longevity due to hydrogel degradation
recordings
Limited capability for high-throughput screening applications
Porous scaffolds[330–342,346,351,352,358] Long-term culture capability of larger scale tissue models Challenging to scale-up for high-throughput screening applications
Greater control of tissue mechanics, particularly useful Matrix needs to provide appropriate cellular microenvironment
for injury models
Suitable for noninvasive imaging and multi-tissue constructs

Adv. Funct. Mater. 2020, 1909146 1909146  (8 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

(CAD) software are used to create photomasks which are used fluidic model of the BBB.[209] These systems demonstrated
to generate a master pattern etched into a photosensitive mate- cell-type specific markers, endothelial barrier, and selective
rial (photoresist). This master pattern is then used as a mold permeability. They also showed neural functionality and could
for casting liquid PDMS, curing the polymer, and peeling the be used in the future as a platform for assessing therapeutic
resultant replica with a negative relief from the master. When effects on neurons as well as BBB permeability.
oxidized in oxygen plasma, the surface of the replica PDMS As in murine models, the contribution of multiple cell
can be sealed using other materials, typically PDMS (patterned types (endothelial cells, pericytes, smooth muscle cells, astro-
or slabs) or glass.[204] This soft lithography approach creates cytes, microglia, and neurons) is critical for BBB formation.
hollow channels for the perfusion of fluids, with applications in The increased availability of hiPSCs, including differentiation
controlling fluid flow and cell patterning in biological systems, protocols for the various cell types, has facilitated the develop-
which offer advantages as compared to traditional 2D cultures. ment of human cell-based models. This has been accomplished
These devices can also be integrated with microsensors to pro- by interfacing human brain-derived microvascular endothelial
vide feedback on culture conditions, including tissue barrier cells (BMECs) with primary human astrocytes and pericytes to
functions. To date, microfluidic brain-on-a-chip systems have provide critical intercellular communication for the establish-
been designed to model the BBB, neural differentiation and ment and maintenance of the high level barrier function of the
brain development, and brain cancer.[205] in vivo human BBB.[210] The further incorporation of neurons,
differentiated from hiPSCs, into a microfluidic BBB system
provides a mechanism for investigating drug effects on neu-
4.2. Blood–Brain Barrier-on-a-Chip ronal function in addition to drug permeability and effects on
barrier function.[211] More recent efforts have sought to create
Barrier functions at tissue interfaces can be modeled and mod- a fully iPSC-derived BBB model using hiPSC-derived BMECs
ulated by seeding cells in different compartments or channels and neural cells (neurons and astrocytes) that can be personal-
of a microfluidic chip. By incorporating a porous membrane ized to detect individual variability in BBB function.[212]
between two microfluidic compartments and seeding with Importantly, these systems can be used to examine the
the appropriate cell types, the microfluidic components can delivery of therapeutics across the BBB. By co-culturing BMECs
be used to control barrier function biochemically or mechani- derived from hiPSCs with rat primary astrocytes, the drug
cally through exposure to fluid shear stress. The fidelity of the permeability of large molecules (fluorescein isothiocyanate
in vitro model to in vivo physiology is governed by the ability (FITC)-dextrans) and model drugs (caffeine, cimetidine,
to replicate key characteristics of the BBB including tight junc- and doxorubicin) were measured and correlated with in vivo
tions between endothelial cells, co-culture with astrocytes, and data.[213] Recent efforts have focused on developing high-
integration of fluid shear stress as well as functional outcomes throughput microfluidic platforms using human immortalized
such as high electrical resistance and selective permeability.[206] cell lines of brain microvascular endothelial cells, pericytes, and
Early models of the BBB used murine cells such as mouse astrocytes suitable for drug screening applications.[214] These
endothelial and astrocyte cell lines seeded on both sides of a systems represent a promising alternative to current options
porous membrane to establish a micro-BBB that mimics the for preclinical studies such as in vitro rodent or human cell cul-
cerebrovascular environment including expression of tight ture or in vivo animal models.[215] In fact, microfluidic systems
junction component zonula occludens-1 (ZO-1) and increased mimicking the BBB are now being used to probe disease mech-
transendothelial electrical resistance (TEER) in response to fluid anisms, such as the role of BBB dysfunction in AD and may
flow. Experimental measures such as a transient drop in TEER be a platform for screening new drugs.[216] The development
in response to histamine exposure as well as selective permea- of microfluidic platforms, combined with the advancement in
bility to fluorescent tracers as a function of molecular size were stem cell technology, will continue to improve understanding
used to evaluate barrier function.[206] In another model system, of the BBB and neurodegenerative disorders, with the ultimate
tight junction molecules were shown to be upregulated by cul- goal of discovering effective therapies.[217]
turing an immortalized rat brain endothelial cell line (RBE4)
with a perfusate of astrocyte conditioned medium in a synthetic
BBB model.[207] 4.3. Platforms for Enhancing Brain Organoid and Neural
While these studies demonstrate the utility of murine cell Spheroid Culture
models of the BBB, they do not wholly recapitulate human
biology. Similar results were observed using human cell lines, Engineering tools and techniques have similarly been used to
as a microfluidic system was used to study the barrier function overcome the limitations of brain organoid and neural sphe-
of immortalized human brain endothelial cells (hCMEC/D3), roid cultures (e.g., reproducibility, heterogeneity, and necrosis).
demonstrating that shear stress increased TEER, while the In contrast to traditional well plates, these systems offer
addition of tumor necrosis factor alpha (TNF-α) had the oppo- spatiotemporal control of culture conditions, which can be
site effect.[208] This further demonstrated the importance of a used to enhance cell viability and to direct neural tissue engi-
shear force component in mimicking barrier function in BBB neering.[218] With the help of microfluidic and microwell arrays,
models, particularly over prior approaches using human cells researchers have uncovered mechanisms of neural differentia-
that were largely focused on static transwells. A hybrid model tion, brain development, and neurological disease. Early studies
using rat cortical neurons and astrocytes, along with human used microfluidic systems to maintain a gradient of cytokines
endothelial cells, was used to form a 3D neurovascular micro- to control differentiation of neural progenitors derived from

Adv. Funct. Mater. 2020, 1909146 1909146  (9 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

human embryonic stem cells into neurons.[219] More recent 4.4. Chip-Based Models of Neurodegenerative Disease
work has focused on immobilizing EBs within an ECM (e.g.,
Matrigel) in a microfluidic channel and cultured under static While the use of brain organoid-on-a-chip systems for studying
or perfused conditions to induce differentiation into brain orga- neurodegenerative diseases is in the early stages of develop-
noids.[220] These systems have been used to model neurode- ment, other neural cultures have used microfluidic systems to
velopmental disorders such as fetal brain dysfunction due to help model Alzheimer’s and other CNS diseases due to their
prenatal nicotine exposure, demonstrating premature neuronal ability to control fluid delivery, incorporate micro-chambers for
differentiation and disrupted cortical development as a function protein accumulation, and establish gradients to study chemo-
of nicotine exposure.[160] Other studies of brain development taxis.[231] Compared to traditional 2D cultures, these are more
have used microfabricated compartments to model the physics representative systems for studying specific pathologies of
of brain wrinkling during the development of human brain neurodegenerative diseases such as propagation of Aβ or tau
organoids.[221] protein through neuronal networks, chemotaxis of microglia,
In addition to using microfluidic chips to control the stem and/or synapse activity. Compartmentalized microchannels
cell environment, microwell arrays have been engineered to con- separating soma from axons have demonstrated early-stage
trol sphere size and shape and aid in homogenous formation of AD pathological events in response to Aβ application such as
3D neural spheroids tissues. Neural spheroids derived from cell transmission of Aβ to cell bodies via axons,[232] impaired axonal
lines, primary rodent brain cells, or human stem cell-derived transport of brain-derived neurotrophic factor,[233] or synaptic
neural cells have been used to recapitulate key features of the dysfunction in a cortico-hippocampal neuronal network.[234]
brain such as cellular microenvironment and diversity.[222–225] In addition to Aβ transport, microfluidic platforms can be
However, because typical 3D neural spheroid fabrication used to study other AD pathologies such as the accumulation
methods such as hanging drop or spinner flask culture do not of microglia near Aβ plaques by observing directional migra-
provide adequate size control and reproducibility, nonadhesive tion of human microglial cells in response to concentrations of
molds, made of agarose or PDMS, or ultra-low attachment monomer and oligomers.[235]
plates are used to allow cellular self-assembly into spheroids Mechanisms behind other hallmarks of AD pathology, such
in a controlled manner.[226,227] Using a micromold technique to as the formation of tau protein aggregates, have similarly been
control size and composition, spheroids prepared from primary studied using microfluidic systems. As with Aβ, a microfluidic
rat cortical cells demonstrated formation of synaptic connec- device with separate somatodendritic and axonal compartments
tions between neurons, electrical activity, brain-native ECM, and was used to demonstrate that human wild-type tau protein
brain-like mechanical properties.[227] Similarly, neurospheres could be transferred between primary rat neuronal cells.[236]
comprised of rat cortical neurons cultured in microwell arrays Using a tripartite microfluidic chamber device, it was fur-
demonstrated controllable size distribution, enhanced viability, ther shown that tau aggregate formation could be induced in
neural architecture, and network formation, as well as decreased downstream neurons through cell to cell propagation as well
cell viability upon exposure to Aβ, a hallmark of AD.[228] Addi- as through extracellular medium.[237] More recently, 3D tissue
tional physiological relevance was achieved by using a concave models have been developed that incorporate multiple human
microarray system and osmotic pump to culture neurospheroids cell types (neurons, astrocytes, and microglia) and hallmarks
under continuous low flow rates to mimic the interstitial fluid of AD (Aβ aggregation, phosphorylated tau accumulation, and
responsible for delivering nutrients and clearing waste in brain glial cell activation), thereby offering a more complete represen-
tissue in vivo. In addition to investigating flow effects in static tation of neurodegeneration and neuroinflammation in AD.[238]
versus dynamic cultures on neural differentiation and network By incorporating many of the advances described in prior work,
formation, neurospheroids were cultured with Aβ to examine this microfluidic 3D triculture advances understanding of path-
neurotoxicity in an in vitro model of AD.[229] ogenic mechanisms, particularly as it relates to glial activation
One of the critical remaining technical obstacles for orga- and inflammatory processes.
noid and spheroid cultures remains overcoming limitations on Similar approaches have also been applied to the study of
the distribution of oxygen and nutrients as organoids develop other neurodegenerative disorders such as Parkinson’s disease by
into larger and more mature tissues. The high cell density, in examining the role of a-synuclein spreading and aggregation as
combination with the absence of active perfusion, can lead to well as neuroinflammation in disease pathology. A microfluidic
a necrotic core in larger (>500  µm) spheroids or organoids. system was used to study intercellular communication between
Miniaturized spinning bioreactors were 3D printed and used human neuroglioma cells and murine microglial cells by control-
to optimize forebrain organoid cultures derived from human ling the fluid exchange of soluble factors between two isolated
iPSCs.[109] These systems provided enhanced cell viability com- cell culture chambers.[239] More recent efforts have focused on
pared to stationary cultures and demonstrated cortical neuro- using patient-derived iPSCs to generate midbrain-specific dopa-
genesis. Exposure of the cultured organoids to Zika virus led minergic neurons to study PD in 3D microfluidic devices. 3D
to characteristic features of microencephaly, offering potential cultures of neurons carrying the LRRK2-G2019S mutation dem-
utility of the bioreactor platform for disease modeling of brain onstrated time-dependent dopaminergic degeneration as well as
development, with subsequent compound testing during drug the preceding mitochondrial abnormalities. Interestingly, treat-
screening.[109] Another future option is to implement microflu- ment with LRRK2 inhibitor 2 rescued dopaminergic phenotype,
idic perfusion networks to drive fluid flow or to grow organoids suggesting that this system may be useful for drug screening
directly onto microfluidic channels seeded with endothelial purposes.[240] High-content imaging systems utilizing auto-
cells to mimic vasculature.[230] mated fluorescent microscopy and analysis algorithms allow for

Adv. Funct. Mater. 2020, 1909146 1909146  (10 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

real-time assessment of these cultures, with other recent efforts as compared to traditional monolayer cultures due to their
aimed at constructing fully automated workstations for differen- ability to recapitulate drug pharmacokinetics in a 3D environ-
tiating PD patient-derived stem cells into dopaminergic neurons, ment. Brain cancer spheroids, cultured from glioblastoma cells
with cultures maintained for over 100 days and monitored for within microwells in a 3D chip, were treated with anticancer
electrophysiological activity in real-time.[241] drugs, pitavastatin and irinotecan, as a combinatorial treat-
ment at varied concentrations in a high-throughput manner to
determine the optimal treatment to reduce cell viability.[250] A
4.5. Brain Cancer-on-a-Chip modified version of this chip was able to culture primary GBM
cancer cells derived from primary human brain tumors and
Microfluidic technologies have been used in brain cancer treat with combinations of chemotherapy drugs, temozolmide
research for diagnostic applications such as circulating tumor and bevacizumab, to determine the optimal concentrations
cell isolation and biomarker detection, as well as to create in vitro for personalized medicine.[251] Importantly, since these micro-
models to study drug efficacy and tumor progression.[242] As chips are prepared from a polymer that blocks nonspecific
with models of neurodevelopment or neurodegeneration, brain protein adsorption, poly(ethylene) glycol diacrylate (PEGDA),
cancer-on-a-chip models rely on precise control of cellular micro- this system avoids the pitfalls of PDMS-based microfluidics
environment and capability for real-time outputs (e.g., high- which can adsorb hydrophobic molecules, an important con-
resolution imaging, biochemical detection) to understand disease sideration for drug screening applications.[250,251] In addition to
pathology. To date, researchers have focused on controlling vari- drug efficacy, other personalized approaches to drug screening
ations in ECM composition and mechanics, vascular integration, have included predicting drug resistivity of GBM cell lines by
and cellular interactions as a means of studying glioblastoma measuring mRNA markers of temozolmide resistance[252] as
multiforme (GBM) tumor progression. As an example, single cell well as determining the drug sensitivity of live slice cultures to
microchips can be used to quantitate the distance dependence chemotherapies in a dose-dependent manner.[253]
of signaling between two GBM cancer cells, providing insight
into the cell–cell interactions that may govern the diffuse nature
of GBM tumor architecture.[243] Tumor progression can be pre- 4.6. Technology and Systems Integration in Microfluidic Platforms
dicted using a microfluidic platform for single-cell proteomics of
human glioblastoma cell lines and brain tumor specimens.[244] The development of advanced in vitro 3D models of human
For understanding tumor metastasis, microfabricated polyacryla- brain physiology and disease through brain-on-a-chip, tissue
mide channels of defined wall stiffness and width were used to engineering or other approaches, has necessitated the integra-
study the effects of matrix composition and geometry on tumor tion of technologies to monitor real-time changes in cellular
cell migration, with narrow channels enhancing migration.[245] microenvironment and activity, particularly in response to
Other microfluidic cell culture devices have been developed to pharmaceutical intervention. As outlined in many of the sys-
investigate the role of hypoxia in maintaining a stem cell niche tems described above, interfacing microfluidic platforms with
in glioblastoma spheroids[246] or to examine clinical hallmarks of optical, electrical, and biochemical capabilities is critical to
glioblastoma such as the formation of migratory hypercellular identifying biological changes in real time. Due to the optical
regions induced by oxygen and nutrient depletion by control- clarity of PDMS and other polymers used in microfluidic sys-
ling the flow of medium using microfluidics.[247] Each of these tems, conventional microscopy techniques (bright field, fluo-
systems demonstrate the utility of 3D microfluidic systems by rescence, phase contrast) have been used to observe cells and
providing insights due to the enhanced control over cell environ- tissues in these devices. As microfluidic devices have trended
ment and experimental variables that are not otherwise achiev- toward thicker 3D tissues, the emergence of scanning-based
able using standard 2D cultures. imaging techniques such as confocal microscopy have allowed
Recent microfluidic models have examined the role of for increased depth resolution.[254] Future integration of light
vascular cells in mediating glioma stem cell invasion. Using sheet microscopy in microdevices may allow for even greater
a microchannel co-culture system of human endothelial spatial and temporal resolution, with lower phototoxicity than
cells and glioma stem cells (GSC) separated by a collagen confocal microscopy.[255] While standard microscopes are still
gel, migration of the glioma stem cells was promoted by the most frequently used, continued development of compact
endothelial cells through the genetic upregulation of cell- microscope-on-a-chip imaging systems will reduce costs and
ECM adhesion-associated proteins and down-regulation of increase capabilities for automation and portability.[256] Future
cell–cell adhesion proteins.[248] A more representative 3D integration of smartphones, utilizing their high-resolution com-
microfluidic model of GSC and vascular cell interactions was plementary metal oxide semiconductor (CMOS) cameras, offers
established with a microvascular network in a fibrin hydrogel another alternative for optical biosensors,[257] with proof-of-con-
prior to introducing patient-derived GSCs in Matrigel to the cept already demonstrated in a microfluidic chip system.[258]
microfluidic device. Using a receptor antagonist, this system For neurological systems, the ability to monitor electrical
demonstrated enhanced migration of GSCs in the presence signaling in real-time is critically important for understanding
of endothelial cells via the chemokine (CXC motif) ligand information processing in neural circuits. For microfluidic sys-
12/chemokine (CXC motif) receptor 4 axis (CXCL12-CXCR4) tems, this is typically achieved through the integration of multi-
signaling pathway.[249] electrode arrays (MEA) that are capable of recording electrical
Brain cancer-on-a-chip platforms are often utilized as more activity with high spatial and temporal resolution. Because
representative model systems for preclinical drug screening these are recorded noninvasively, repeated measurements

Adv. Funct. Mater. 2020, 1909146 1909146  (11 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

can be achieved over time in either brain slices or tissue cul- 5. Materials-Based Approaches to 3D Brain Models
ture preparations.[259] Based on the design of the microfluidic
system and MEA, research can range from investigating intra- Whether in a chip-based or stand-alone system, the choice
cellular dynamics in a single axon[260] to propagation between of materials for a 3D brain model is multifaceted, with con-
neuronal sub-populations (e.g., cortical-thalmic) in a co-culture siderations for biocompatibility, mechanics, porosity, and
system.[261] Although not part of a microfluidic chip, MEAs imaging compatibility, among the key concerns. Neural cells
have also shown utility in drug screening applications such as must be able to proliferate, differentiate, and form viable net-
the evaluation of neurotoxic compounds.[262–265] Importantly, works within materials that also need to provide the appro-
electrical recordings by MEAs can be combined with optical priate porosity, structure, and mechanical support over time.
methods for a more comprehensive assessment of neuronal Additional biophysical cues such as electrical conductivity and
function. As an example, by loading fluorescent calcium sen- architectural patterning are increasingly being integrated into
sitive dye (Fluo-4) in the sample medium, simultaneous material designs to guide cell growth, development, and organi-
recordings of extracellular calcium and electrical traces can be zation.[280] In addition to support structure, materials must also
resolved to show a propagating wave in response to electrical incorporate domains for cell attachment as well as appropriate
or chemical stimulation of murine brain slices.[266] Though bioactive signaling molecules (e.g., immobilized growth fac-
not explicitly for brain organoids, the recent development of a tors or chemical functional groups). By carefully designing and
noninvasive, automated platform for physical, biochemical, and engineering the appropriate biochemical and biophysical cues,
optical monitoring represents the latest achievement in real- researchers can direct cellular behavior by recapitulating the
time multisensory feedback in a microfluidic device.[267] natural cellular microenvironment in a 3D brain model.[281] For
In addition to advances in the hardware for MEA and many researchers, particularly in the organoid field, Matrigel
imaging techniques, protein and genetic engineering is the chosen support material as it offers a complex mixture
approaches, such as genetically encoded calcium indicators of structural, functional, and signaling molecules suitable for
(GECIs) or optogenetic techniques, are increasingly being used 3D culture, but remains a poorly defined and heterogenous
as alternative methods to voltage-sensitive fluorescent dyes to mixture of ECM proteins that is subject to batch-to-batch vari-
observe, analyze, and control neuronal activity. The develop- ability.[282] To address these deficiencies, a variety of material-
ment of a high affinity, green fluorescent protein-based calcium based approaches have been used to bioengineer 3D scaffolds
(Ca2+) probe (GCaMP) allows for chronic imaging and quanti- using a bottom-up approach to support in vitro neural tissue
fication of changes in intracellular calcium concentration as a development and disease modeling.
function of firing action potentials and synaptic input.[268,269] By
transfecting neurons with GCaMP and culturing them in 3D
hydrogel matrices, multi-planar images provide spatial speci- 5.1. Natural Hydrogel Systems
ficity and the functional activity of the developed 3D neural net-
works, particularly in response to drug input.[270,271] Additional Hydrogels are commonly used in bioengineered neuronal cul-
fluorescent protein-based probes such as red GECIs (RCaMP) ture systems as their chemical and mechanical properties can
provide the option of using different colors within the same cell be tailored to resemble those naturally occurring in brain.[283]
or to distinguish cell populations (e.g., neurons and astrocytes) Derived from natural or synthetic polymers, they are cross-
and may be integrated with optogenetic approaches.[272] linked to form hydrated networks of tunable mechanical stiff-
The emergence of optogenetics, starting from the use of ness and may be functionalized with bioactive components.[283]
a light-sensitive ion channel protein (Channelrhodopsin-2, In contrast to Matrigel systems, hydrogels prepared from fully
ChR2), as a means of optically controlling the activity of neu- defined source materials can be formulated with specific con-
ronal populations has provided researchers with another tool to trol over ECM components and signaling cues, using in vivo
activate or inhibit the activity of specific neurons noninvasively conditions as a guide. Because brain ECM is composed of
in culture with high temporal and spatial resolution.[273–275] a mixture of proteoglycans, hyaluronic acid, tenascins, col-
While the majority of development is focused on in vivo appli- lagens, fibronectin, vitronectin, and laminin,[284] combina-
cations, such as neural probes for brain stimulation, microflu- tions of these materials have been used as a starting point to
idic systems have used optogenetics and calcium imaging to engineer hydrogel scaffolds for neural tissue engineering.[282]
observe synaptic communication in vitro.[276] 3D in vitro tissue Importantly, hydrogels must be formulated to mimic the
models have demonstrated the ability to record changes in mechanical properties of the brain (elastic modulus of < 1 kPa),
fluorescent intensity upon photostimulation of neural cells that providing an appropriately soft matrix to direct neuronal cell
are transfected with both a ChR2 and a GECI and cultured in a behavior.[285] This is typically achieved by preparing hydrogels
collagen gel on a CMOS chip.[277] As optogenetic tools continue that are cross-linked with sufficiently low density to achieve this
to be developed, additional methods of cellular control and/or mechanical stiffness while also allowing for exchange of oxygen
control of small molecule delivery in 3D systems are available and soluble factors to support neuronal culture.
for researchers to study different mechanisms of neuromodula- Due to its wide availability, biocompatibility, and biodegrada-
tion.[278] As a key example, the ability to combine optogenetic bility, as well as its well-documented historical use as a hydrogel
actuators (channelrhodopsins) with archaerhodopsin-based scaffold for tissue engineering, the most commonly used nat-
voltage indicators in hiPSC-derived neurons allows for nonin- ural biomaterial for neural cell culture applications is collagen
vasive, all-optical electrophysiology that is well-suited for high- type I. It has been demonstrated that rodent-derived neural
throughput drug screening.[279] stem and progenitor cells (NSPCs) in type I collagen gels can

Adv. Funct. Mater. 2020, 1909146 1909146  (12 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

expand and differentiate into neurons with functional synapse binding domains has been to mix in collagen fibrils, with this
and neuronal circuit formation.[286] These gels have also shown collagen/alginate blended hydrogel used to culture neurons
the ability to support the differentiation into multiple cell types generated from human iPSCs.[301] Other researchers have used
(neurons, astrocytes, and oligodendrocytes) under optimal con- defined mixtures of nutrients in agarose and alginate hydrogels
ditions of cell density and collagen concentration.[287] In order to generate 3D multi-layered scaffolds that mimic cortical layers
to further increase relevancy to native brain tissue, additional using primary rat cortical neurons in a microfluidic device.[302]
natural polymers such as hyaluronan have been combined with Though the vast majority of neuronal tissue engineering has
collagen to enhance neuronal differentiation of NSPCs.[288] been achieved using rodent-derived neurons, combinatorial
Similarly, the incorporation of soluble growth factors in engi- studies on matrix composition and architecture have sought
neered hydrogels have been used to direct cell proliferation, to provide a more comprehensive understanding of how cel-
migration, and fate. Using a layer-by-layer bioprinting approach, lular microenvironment (ECM and signaling molecules) can
collagen-fibrin composite hydrogels were used to promote the influence cell fate in 3D cultures. This includes efforts to use
migration of neural stem cells toward hydrogel layers containing these parameters to direct human stem cell differentiation
vascular endothelial growth factor (VEGF).[289] Combinations of toward neural and glial lineages in natural polymers such as
neurotrophic factors such as nerve growth factor (NGF), glial Matrigel, collagen type I, gelatin, or HA.[303] As the field pro-
cell-line derived neurotrophic factor, and ciliary neurotrophic gresses toward high-throughput, human cell-based neuronal
factor have been explored in dorsal root ganglion (DRG)-seeded models for developmental/disease biology, regenerative medi-
collagen gels to optimize neurite outgrowth.[290] Mechanical cine, or drug/toxicity screening applications, researchers aim
cues such as gradients of stiffness in 3D collagen gels or bio- to overcome limitations on reproducibility and culture duration
acoustic levitation in fibrin gels have also been utilized to drive that is inherent for natural materials. One approach is the use
neurite growth or generate multilayered 3D brain-like con- of novel biomaterials such as silk fibroin protein to generate
structs from hESC-derived neuroprogenitors, respectively.[291,292] hydrogels with tunable stiffness in the range of native neural
Due to the prevalence of hyaluronic acid (HA) in native tissues, capability for functionalization with growth factors or
brain tissue, this polymer has been extensively studied as a sub- ECM molecules, as well as the ability to maintain structural
strate for central neural tissue engineering.[293] However, since integrity for much longer duration (months to years) than
unmodified HA does not support cell attachment, HA hydro- typi­cal hydrogel materials.[304] Others approaches center around
gels must be modified with cell adhesion molecules or blended the development of synthetic materials, including polymers and
with other ECM components such as collagen.[294] As an self-assembling peptides, to address these concerns.
example, a Nogo-66 receptor antagonist was used to modify HA
hydrogels to improve the neuronal adherence and survival of
DRGs in vitro as well as induce neurite outgrowth.[295] HA-poly- 5.2. Synthetic Hydrogel Systems
D-lysine (PDL) copolymer hydrogels have also been used to
control neuronal cell adhesion and network formation based on In contrast to naturally derived polymers which offer physi-
HA:PDL ratios.[296] Similarly, ECM components such as laminin ological relevance but also greater variability in chemical com-
or fibronectin- or laminin-derived peptides (RGD, IKVAV) have position, synthetic hydrogels derived from polymeric materials
been used to facilitate cell attachment to HA.[294] A combination such as polyethylene glycol (PEG) and methacrylates, among
of HA, collagen, and laminin was used as a mimic of natural others, offer the advantage of tight control over physical and
ECM composition to support the co-culture of Schwann cells mechanical properties based on chemistry and polymeriza-
and neurons in 3D hydrogels for 2 weeks, offering a potential tion.[283] However, these hydrogels are generally biologically
platform for studying mechanisms of interactions between the inert and with minimal cell adhesion and protein adsorption
two cell types with applications in nerve regeneration.[297] Inter- to support tissue formation.[305] This requires chemical modifi-
estingly, as with collagen hydrogels, mechanical properties of cations to the polymer hydrogels for neural tissue engineering
HA gels influenced neural progenitor cell (NPC) differentiation applications that are similar to those used for polysaccharide-
toward neuronal subtype in softer hydrogels versus an astro- derived hydrogels (e.g., RGD, IKVAV) or mixtures with other
cytic subtype in stiffer hydrogels.[298] ECM proteins (e.g., collagen). As with chemically defined nat-
In addition to HA, other naturally derived polysaccharides ural hydrogels, the use of synthetic polymers is an alternative to
such as alginate, agarose, chitosan, and composites thereof Matrigel, replacing the inherent variability of a poorly defined
have been used as substrates for 3D neural tissue engineering. mixture of bioactive components with a well-defined and con-
As with HA, these polymers are inert biomaterials that can be trolled matrix for neural tissue engineering applications.
tailored for 3D neural cell culture by varying composition and To date, PEG is the most extensively used synthetic polymer
crosslink density. These polymers are similarly nonadhesive for for neural cell culture in 3D hydrogels. Preliminary studies
cultured neurons and must incorporate cell attachment factors, incorporated rat neurons in photopolymerizable PEG hydrogels,
although there is some evidence that unmodified soft gels can modified with poly(lactic acid) and poly(glycolic acid) to control
support neurite growth.[299] As an example, alginate hydrogels degradation rate and supplemented with basic fibroblast growth
functionalized with cell-specific attachment ligands (laminin, factor 2 (bFGF-2) to support cell growth.[306] These tissues dem-
IKVAV, or RGD) were used to study co-cultures of neurons, onstrated neural tissue development and functionality, with the
glia, and endothelial cells, observing neural functions such as time-scale of neurite outgrowth a function of gel degradation rate.
gap junction formation, synaptic vesicle release, and electro- In another study, peripheral nerve regeneration was enhanced
physiological activity.[300] Another approach to incorporating cell by covalently attaching cell adhesion ligands (RGD, IKVAV,

Adv. Funct. Mater. 2020, 1909146 1909146  (13 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

and YIGSR) to PEG hydrogels.[307] Adhesion ligands conjugated offers new possibilities for building 3D neural tissues in vitro
to PEG hydrogels have similarly been used to study human using a bottom-up approach. Using hydrogel-based bioinks to
iPSC-derived NPCs attachment and migration induced by precisely arrange cells in 3D, heterogenous tissues can be com-
astrocytes.[308] Recent advancement in PEG hydrogels has been posed with the appropriate physiological cues and resultant
achieved by incorporating various functional groups, including functionality that is similar to that of native brain tissue. This
polymer-peptide conjugates, for cell attachment, signaling, approach further improves reproducibility in model develop-
and cleavage sites to tune physical and mechanical properties. ment and potential for automation, critical elements for down-
Using star-shaped PEG (starPEG) and heparin peptide conju- stream high-throughput screening applications.[222]
gates, hydrogel compositions were optimized to embed human Many of the hydrogel materials described above have been
vascular endothelial cells and DRGs and stimulated to control investigated as potential bioinks for neural tissue engineering
cell phenotype (angiogenic or neurite outgrowth).[309] applications, with the choice dependent on a number of con-
Synthetic polymers have been used to modify natural siderations including the printing technique (e.g., inkjet,
polymers, preserving the beneficial properties of the natural extrusion-based, laser-guided), cytocompatibility, curing time,
material while offering additional functionality. One notable and mechanical properties, among others.[321,322] In one of the
example is the synthesis of methacrylamide chitosan (MAC), a first examples of bioprinting neurons, inkjet printing was used
synthetic polymer that can be modified with peptides for cell to pattern rat hippocampal and cortical cells in a single layer,
adhesion and neurite outgrowth, formed into a transparent with these cells demonstrating electrophysiological activity.
gel via chemical crosslinking, and seeded on the surface with Researchers in this study used the same technique to fabri-
rat neurons.[310] This polymer has subsequently been modified cate multi-layer constructs of human NT2 neurons and fibrin
to enhance oxygen diffusion within the hydrogel through the gels via a layer-by-layer printing approach to create 3D neural
incorporation of D-mannitol[311] or conjugation of perfluorocar- sheets.[323] A similar approach using multilayered collagen type I
bons to MAC[312] to enhance the neuronal differentiation of rat gels and rat embryonic neurons and astrocytes was used to
neural stem cells (NSCs)/progenitor cells (NPCs) in 3D. Meth- form multicellular constructs with defined spatial patterning
acrylate has also been used to modify HA, with researchers to examine cell-to-cell interactions.[324] Layer-by-layer inkjet
using this material to create layered 3D hydrogels of variable printing has also been used to study cell migration through the
mechanical properties to study the differentiation and migra- controlled spatial distribution of VEGF-containing fibrin gel
tion of human iPSC-derived NPCs.[308] and murine NSCs within a 3D collagen scaffold.[289] Subsequent
Synthetic, self-assembling peptides (SAPs) have been work has focused on exploring tunable polymers as inks for
developed as a new class of biologically inspired materials 3D bioprinting such as thermoresponsive polyurethane hydro-
for neuronal culture. Oligopeptides of varied sequences of gels[325] or stereolithography-based printing of functionalized
arginine-alanine-aspartate (RAD) are able to self-assemble materials such as dopamine-modified gelatin methacrylate[326]
under physiological conditions and support neuronal cell to improve NSC proliferation and differentiation.
attachment, neurite outgrowth, and synapse formation,[313] with Improvements in printing techniques and bioink material
the most commonly used peptide, RADA16-I, commercially properties have supported the fabrication of more complex
available as PuraMatrix. Neural stem cell proliferation, migra- structures, such as cortical layers, that mimic the native brain.
tion, and differentiation may be further promoted by modifying Through the use of a bioink (gellan gum-RGD) that forms a
these SAPs with a functional cell adhesion motif (RADA16- porous structure suitable for oxygen and nutrient exchange,
IKVAV) in a 3D hydrogel scaffold using both rodent and human primary rat neural cells were printed in a layered, brain-like
cells.[314–316] Despite these successes, these formulations suffer cortical construct.[327] More recently, human NSCs and iPSCs
from low pH impacting cell viability. To address this issue, in a polysaccharide-based (alginate, carboxymethyl-chitosan,
oppositely charged SAPs with different cell adhesion epitopes agarose) bioink was used to print mini-tissues that differentiate
(RADA16-IKVAV and RADA16-RGD) were combined to form a into neurons, form neural networks, and demonstrate electrical
3D hydrogel at neutral pH that was capable of encapsulating activity. Using an extrusion-based printing technique, cells are
NPCs/NSCs and differentiating into neurons and astrocytes.[317] co-printed with the bioink before gelation to encapsulate them
Because SAPs often suffer from poor mechanical properties, in a 3D matrix for in situ expansion and differentiation into
linear or branched (LDLK)3 SAPs were recently used as an alter- functional neurons and neuroglia.[328,329]
native to form hydrogels of controlled mechanical properties,
demonstrating that neuronal differentiation is enhanced when
scaffold stiffness is comparable to that of the native brain.[318,319] 5.4. Solid Porous Scaffolds
These constructs have also demonstrated the ability to sup-
port in vitro, serum-free differentiation of human NSCs into While hydrogel-based techniques offer the benefit of soft
GABAergic, glutamatergic, and cholinergic neurons that give mechanical properties for neurite outgrowth, they generally
rise to the formation of electrically active networks.[320] lack the ability for long-term culture (greater than 1 month) due
to fast degradation rates, uncontrolled contraction, or the pres-
ence of necrotic cores. As an alternative, solid porous scaffolds
5.3. Bioprinting prepared from synthetic or natural polymers offer mechanical
and morphological stability and pore interconnectivity to facili-
The emergence of bioprinting as a technique to control the spa- tate nutrient transport required in long-term neural tissue
tial distribution of cells, biomaterials, and signaling molecules models. Furthermore, these scaffolds may be modified with

Adv. Funct. Mater. 2020, 1909146 1909146  (14 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

topographical cues for cell attachment and directed neurite activity. The composite system (silk and collagen type I) was
outgrowth.[222] further used as a model of TBI, monitoring changes in cellular
As an example, researchers have used polystyrene as a syn- damage, biochemical signaling, and electrical activity following
thetic polymer for generating porous scaffolds for 3D culture a controlled impact.[341] These versatile systems have also
of human neural progenitor cells. Scaffolds were prepared demonstrated utility as engineered neural tissue implants for
by a gaseous salt leaching method where polystyrene is dis- replacing large brain cavities.[342]
solved in chloroform, mixed with salt particles of known size,
and cast into wells before curing in an oven.[330] When using
these scaffolds to culture human neural progenitors, the cells 5.5. Materials-Based Models of Neurological Disorders,
more accurately emulated the gene expression profiles of in Damage, and Dysfunction
vivo surrogates (3D neurospheres) in comparison to standard
2D culture.[331] Porous polystyrene scaffolds cultured with pri- In addition to supporting 3D cultures to study neurite out-
mary rodent neural cells also demonstrated more representa- growth and neural network formation, materials-based
tive calcium channel functionality versus standard 2D cultures approaches have been utilized to create models of neurological
in comparison to freshly dissected intact tissues.[332] More disease. To date, work has primarily focused on the develop-
recently, graphene foam has been pursued as a biocompatible ment of 3D in vitro models of AD and glioblastoma as well
and conductive scaffold for 3D neural culture that represents as platforms for studying TBI or neurotoxicity. By controlling
an alternative to the inert, nonbiodegradable polymers such material properties as well as cell type and genetic composition,
as polystyrene. These scaffolds, prepared via a chemical vapor particularly using human iPSCs or neural progenitors, key fea-
deposition method, showed that they can enhance NSC differ- tures of the disease may be recapitulated and used to under-
entiation into neurons/astrocytes as well as serve as a conduc- stand mechanisms and/or to screen therapeutic interventions.
tive platform for electrical stimulation.[333] As an alternative to Initial 3D models of AD have sought to recapitulate the
porous scaffolds, electrospinning techniques have been used pathological hallmarks of the disease, Aβ accumulation, and
to prepare fibrous scaffolds from synthetic polymers that can plaque formation as well as aggregation of phosphorylated tau
be functionalized to support neural cell culture and neurite protein. In one example, fAD mutations in Aβ precursor pro-
outgrowth. However, because standard methods typically pro- tein (APP) and presenilin 1 (PS1) genes were overexpressed in
duce dense fiber meshes that function more as a 2D surface, human NPCs. These cells were cultured in Matrigel as a 3D
recent efforts have used an array of needles as a collector to support matrix, which limited diffusion and accelerated β depo-
form lower density electrospun fiber scaffolds that support 3D sition and tau protein aggregation. This model system recapitu-
culture of human NPCs and their formation of highly inte- lated, for the first time, these key features of AD in a 3D model
grated, functional networks.[334] and was subsequently used to screen beta- or gamma-secretase
Similar to synthetic polymers, multiple methods have been inhibitors to decrease Aβ pathology.[86] In another study, Aβ
used to create porous or fibrous 3D scaffolds from natural aggregates were added to conventional 2D cultures and 3D
polymers. Freeze-drying, in particular, has been commonly cultures of human stem cell-derived neurons in a PuraMatrix
used to form porous scaffolds from biological materials such hydrogel. Cellular response to the Aβ aggregates, as mediated
as collagen or HA. The freeze-drying process can be manipu- by p21-activated kinase, was only observed in the 3D cultures
lated to control the mechanical properties and to determine demonstrating the critical role of the 3D environment in mod-
pore size of collagen-glycosaminoglycan scaffolds.[335,336] eling AD mechanisms.[343] Subsequent 3D models of AD have
Collagen and HA were mixed at different ratios, freeze dried, used collagen or PEG hydrogels to study Aβ-induced toxicity
and cross-linked to tailor the resultant scaffolds to have moduli of neural cell lines or plasticity loss of human neural stem
and porosity similar to those of native brain tissue.[337] This pro- cells, respectively.[344,345] Others have used a composite silk
vides the appropriate conditions to support neural progenitor scaffold-collagen gel system to support long-term (>8 month)
cells by allowing for the exchange of nutrients, the outgrowth 3D human neural network models using iPSCs derived from
of neurites, or the infiltration of cells in in vivo applications.[338] healthy individuals and AD patients, offering a potential plat-
Collagen-based scaffolds can be tailored to include cell adhe- form for studying development, plasticity, and degeneration
sion molecules as in the modification of 3D freeze-dried col- over time.[346] The continued development of 3D models that
lagen matrices with IKVAV and/or RGD peptide to enhance incorporate multiple cell types (neurons, astrocytes, microglia,
the attachment of rat DRGs.[339] Cryogels from gelatin-laminin and oligodendrocytes) will be critical in understanding the
were used as porous scaffolds for differentiating human cord mechanisms of neurodegeneration and neuroinflammation
blood-derived stem cells into artificial neural tissue with mature that are key features of neurodegenerative disorders such as
networks of neurons and glia, suitable for implantation in rat AD or PD.[347]
brain tissue or high-throughput personalized medicine applica- In addition to AD, the development of in vitro 3D brain
tions.[340] Recently, porous silk scaffolds have been used in com- cancer models has received a significant amount of atten-
bination with collagen hydrogels to create functional brain-like tion, with specific focus on materials-based approaches to
cortical tissue, where the compartmentalization of neuronal cell mimic the tumor microenvironment in vitro that drives glio-
bodies and axons recapitulates gray and white matter.[341] This blastoma (GBM) progression. Currently used bioengineered
combination of porous scaffold and hydrogel matrix was able to models are primarily based on the encapsulation of glioblas-
mimic the mechanics of the native rodent brain while allowing toma cells in 3D biomaterial microenvironments that recapitu-
for a robust neural network formation and electrophysiological late the hypoxic conditions observed in vivo and allow for cell

Adv. Funct. Mater. 2020, 1909146 1909146  (15 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

migration within the scaffold.[348] Through the use of natural predictive of chemical safety and reduce the number of drug
ECM-derived materials, primarily HA, researchers are able to failures in clinical trials. Recently, human embryonic stem cell-
mimic the composition and structure of the native brain. In derived neural progenitor cells, endothelial cells, mesenchymal
fact, glioma cell invasion into HA hydrogels has been shown stem cells, and microglia/macrophage precursors formed
to be mediated by controlling the stiffness of HA-methacrylate neural tissue constructs following seeding on top of PEG hydro-
hydrogels and functionalization with RGD peptides for cell gels functionalized with biodegradable peptide cross-links and
adhesion.[349] One of the major drawbacks of using these RGD peptides to promote cell adhesion. Leveraging the repro-
systems is that by increasing polymer concentration or cross- ducible and controllable intracellular interactions achievable via
linking density, biochemical signaling is affected. Others have this method, these constructs were then used for neurotoxicity
sought to independently control matrix stiffness using syn- screening in vitro by treating the tissues with toxic and non-
thetic polymer-based hydrogels such as polyacrylamide or PEG toxic chemicals, collecting RNA sequencing data, and then
to isolate the effects of matrix stiffness and confinement on using machine learning to build a predictive model capable of
human glioma cell migration.[245,350] In addition to hydrogels, correctly predicting the toxicity of additional compounds.[360]
solid porous scaffolds of chitosan and HA have been devel-
oped to study glioblastoma microenvironment, demonstrating
enhanced invasiveness of human GBM cells and chemotherapy 6. Future Perspectives for 3D Brain Models
drug resistance.[351] Using porous silk scaffolds infused with
ECM hydrogels as a brain-mimicking microenvironment, Over the past decade, there has been tremendous progress in
researchers have further showed that pediatric and adult brain the development of 3D models of the human brain. Through
tumor progression is mediated via ECM-dependent alterations technological advances and innovation in the key components
in transcriptomic and metabolic profiles.[352] Additional spatial of engineered tissues, cells, scaffolds, and signaling molecules,
control may be achieved by using bioprinting approaches to otherwise known as the tissue engineering triad, researchers
control the deposition of different cells types as well as ECM can now create models that offer greater fidelity to neurolog-
and signaling molecules in complex 3D models of the tumor ical development and disease. These models offer the ability to
microenvironment.[353] As an example, a 3D brain tumor model better understand human physiology and the potential to iden-
was developed from glioma stem cells bioprinted in a gelatin/ tify therapeutic or regenerative medicine approaches to treat
alginate/fibrinogen hydrogel. Similar to conditions found in disorders of the human brain. As outlined in this review, there
vivo, these cells formed spheroids, remodeled ECM to occupy have been two generally parallel approaches for creating 3D
a larger space, and demonstrated chemotherapy drug resist- brain culture systems, organoid technology, and bioengineering
ance.[354] Spatial control of cells, matrix, signaling molecules, methodologies. The fundamental difference between these
and oxygen content may also be controlled using microfluidic approaches is that organoid technology is largely a cell-driven,
mixing as demonstrated in a HA-gelatin-methacrylate hydrogel bottom-up approach whereas the bioengineering methodology
brain tumor model.[355] More recently, bioprinting techniques is a materials-templating bottom-up approach for generating 3D
have been used to create 3D mini-brains that include GBM brain models. Looking forward, we see tremendous potential
cells as well as macrophages to study cellular interactions and for integrating these two approaches, leveraging the strengths
therapeutic control of those interactions.[356] With these 3D of each. Furthermore, we see potential for integrating other
systems, the goal is to not only understand the mechanisms tools and techniques such as microfluidics and bioreactors for
of tumor progression, but also serve as a platform for preclin- improving cell culture duration or electrode arrays and imaging
ical screening of therapeutics and/or radiotherapy.[357] Using modalities for noninvasive monitoring of the models in real
GBM cell lines seeded in a 3D Matrigel-coated polystyrene time. Lastly, as the potential for 3D models as a screening tool
scaffold, researchers showed that the in vitro cellular response is currently limited by reproducibility issues, we highlight pro-
to radiotherapy in combination with clinically used agents spective solutions for generating high-throughput 3D brain
(temozolomide, bevacizumab, erlotinib) strongly correlated models for biomarker or drug screening applications (Figure 3).
with clinical observations.[358] As described in previous sections, organoids have proven to
In addition to studying neurological disease, recent work has be an effective developmental model of the human brain with
sought to develop preclinical models for studying TBI or neu- additional utility as a tool for studying neurological function
rotoxicity. These models are used correlate clinical observations or modeling neurodegenerative, psychiatric, and infectious
with in vitro results to better understand mechanisms and/or disease. The drawbacks associated with organoid cultures,
to more accurately predict in vivo outcomes than traditional 2D including issues with scalability, reproducibility, heterogeneity,
cultures or animal models. For TBI, collagen gels seeded with and longevity due to necrotic cores, remain active areas of pur-
rat cortical neurons were subjected to compressive deforma- suit in the field. While microfluidic or bioreactor approaches
tion, providing insight on how neuron viability and function is have most commonly been used to improve the reproducibility
dependent on strain magnitude and rate.[359] Others have used of organoid cultures with regard to overall size distribution and
a weight drop method to inflict mechanical damage in a 3D neural cell viability and differentiation, a less explored option
model of cortical brain-like tissue formed from a silk-collagen is the use of materials-based approaches to improving brain
composite scaffold, observing enhanced glutamate release and organoid cultures. In one example, bioengineered materials
transient hyperactivity that is consistent with in vivo results.[341] were used to help developing organoids adopt 3D architectures
For neurotoxicity studies, 3D in vitro models using human cells similar to the native brain as culturing EBs on floating scaffolds
represent an alternative to animal models that may be more of synthetic polymer (PLGA) filaments caused self-organization

Adv. Funct. Mater. 2020, 1909146 1909146  (16 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

Figure 3.  Proposed integration of technology platforms for optimizing 3D tissue systems for modeling human brain physiology and high-throughput
screening. Thicker lines between platforms indicate current areas of investigation whereas thinner lines show areas for further consideration. Gray
boxes identify some of the key stem cell and bioengineering technology parameters to be addressed.

in an elongated morphology and the formation of a polarized to direct neural differentiation. In fact, placing morphogen
cortical plate and radial units.[40] Others have used defined, gradients within organoids has been shown to facilitate tissue
modular, synthetic matrices composed of PEG-based engi- self-organization.[363,364] This could be further optimized
neered hydrogels to understand the role of biochemical and bio- through the patterning of a bioengineered scaffold with growth
physical cues in regulating early neurogenesis in organoid cul- factors known to direct neural progenitor differentiation to
tures derived from mouse ESCs. By controlling the mechanical specific lineages. Cell migration and self-organization within
stiffness, degradability, ECM components, and soluble factors the organoid could similarly be directed by controlling the
of the engineered hydrogels, they were able to identify the spatial organization of cell-adhesive ligands, biochemical cues
specific material properties that promoted neuroepithelial dif- (including gradients), and/or mechanical properties.[362]
ferentiation, apical-basal polarity, and dorsal-ventral patterning, Controlling these aspects of cell–material interactions have
while also generating more homogenous cultures in these been explored using synthetic or biological materials that
defined synthetic matrices in comparison to ESC cultures in otherwise do not possess cell instructive cues, notably HA or
Matrigel.[361] These pioneering studies demonstrate the great silk fibroin, and using the materials as a blank slate to design
potential for integrating organoid and bioengineering tech- cell–matrix interfaces.[365,366] Additional spatial control may
nology to support brain organoid cultures, but with significant be achieved through continued improvement in 3D-printing
insight to be gained by studying bioengineering approaches technology such as the printing of gradients of signaling
used in other in vitro organoid-based tissue models. molecules to direct neural stem cell differentiation.[367]
To date, engineered materials for supporting organoid cul- Other recent efforts using biocompatible electroconductive
tures have mainly focused on solutions for intestinal orga- materials, such as graphene or polypyrrole, offer another
noid tissue models. Through these studies, the importance potential option for guiding the formation of neural networks
of providing a reproducible, chemically-defined matrix with in a 3D system through the use of electrical signaling.[366]
controllable mechanical properties and defined signaling cues This includes combinations of these materials to create novel
for specific stem cell niches has been established. The knowl- matrices for neural tissue engineering applications.[368,369]
edge gained from these studies can be transferred to guide the In each of these examples, the engineered materials would
optimization of materials for 3D brain organoid culture.[362] serve to improve on the commonly used matrices, which are
Among the key goals of an engineered material for this applica- often undefined (e.g., Matrigel) or with limited complexity
tion would be the spatiotemporal delivery of biochemical cues (i.e., build of a single polymer, such as collagen). Such an

Adv. Funct. Mater. 2020, 1909146 1909146  (17 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

approach would offer the potential to improve the consistency approaches to rescuing activity. Another pressing issue in the
of organoid cultures through the use of tunable and reproduc- development of 3D brain models is the use of iPSCs as rep-
ible matrices that also provide the versatility of incorporating resentative models of neurodegenerative disease. iPSCs have
multiple signaling cues (biochemical, mechanical, spatial, been “aged” through genetic and chemical approaches inter-
electrical) as points of control. fering with specific pathways, such as mitochondria or DNA
To address the formation of necrotic cores during 3D brain replication, and by performing direct differentiation of aged
tissue development, a key limitation of organoid cultures, fibroblasts to neurons. However, these approaches remain arti-
engineering approaches may be used to control scaffold archi- ficial and reduce the aging phenotype to a single factor.[377–387]
tecture to allow the proper exchange of oxygen, nutrients, and For each of the technologies described above, one of the
waste products. One approach has been to expose cerebral orga- main goals is to improve the reproducibility of 3D brain
noids in liquid culture to an air interface which augments the models for biomarker or drug screening purposes. In consid-
survival and axonal outgrowth of neurons, presumably through ering systems that will be useful for this application, some
an increased nutrient supply.[370] For a more sophisticated bio- key issues include ease of fabrication and use as well as
engineering approach, porous scaffolds have been used to sup- confounding factors such as nonspecific adsorption of thera-
port 3D neural tissue models, but the continued development peutics to materials used in the system. As an example of a
of new fabrication methods using computer-aided design and reproducible scaffold system, researchers used an automated
3D printing offers unprecedented levels of control for pre- process to incorporate 3D polystyrene scaffolds into standard
paring porous scaffolds.[38,371] Specialized architecture may 96-well plates and were able to create a high-throughput
be used to design suitable implants for regenerative medicine capable platform for culturing neural stem cells and evalu-
purposes in vivo or to support the long-term in vitro culture of ating outputs in growth, gene expression, and functionality via
neural tissue-engineered constructs, which may be critical for a calcium assay.[330,388] Other material-based approaches for
the examination of neurodegenerative diseases such as AD or high-throughput applications will similarly need to demon-
PD. In addition to scaffold porosity, the other commonly used strate the capability for automated production of reproducible
engineering approaches to support chronic cultures is the use scaffolds. Another consideration for drug screening applica-
of microfluidic or bioreactor systems. Bioengineered scaffolds, tions, particularly for microfluidic or chip-based approaches,
particularly hydrogels, are often limited in culture duration due is the nonspecific adsorption of therapeutics. Because hydro-
to concerns over relatively fast degradation rates or restricted phobic molecules adsorb to PDMS, alternative materials such
oxygen/nutrient diffusion for dense cultures. Early bioreactor as silk, polyurethane, or others have been proposed as the
systems such as the rotary wall vessel bioreactor have been starting material for microfluidic systems. As an example, poly­
utilized to grow cortex-like tissue constructs without necrotic urethane elastomers demonstrated similar beneficial optical
cores in collagen gels.[372] More recently, microfluidic or micro- and mechanical properties to PDMS but offered improved
array technologies have been used, but many cultures are still resistance to adsorption of a hydrophobic small molecule
short-term (less than 1 month). We expect the continued devel- dye.[389] Other researchers used PEGDA, a synthetic polymer
opment of bioreactor systems to support longer-term culture that blocks nonspecific protein adsorption, to develop a brain
will help in recapitulating the pathophysiology of neurological cancer chip for high-throughput drug screening. Using this
disorders.[205] Furthermore, given the recent interest in the system to culture glioblastoma cells in a 3D brain cancer
association of brain disease with dysfunction in other physi- model, they were able to examine combinatorial treatments
ological systems (e.g., the gut-brain axis), the incorporation of chemotherapy drugs, pitavastatin and irinotecan.[250] Sub-
of multiple systems in multi-compartment chips or bioreac- sequent systems should seek to achieve a similar ease of use
tors will help identify the impact of peripheral tissues in brain and applicability in developing a commercially and clinically
neurodegeneration.[373] promising high-throughput platform.
The combinatorial nature of neurological function and dys- Another important consideration for commercial and clinical
function has necessitated innovations in the ability to model applicability is the integration of technological innovations in
interactions between different cell subtypes, brain regions, and electrical or optical interfaces for monitoring 3D brain tissue
other tissues. Brain tissue models with higher complexity have models in real-time for high-throughput screening applica-
been achieved by fusing different types of spheroids together tions. While automated fluorescence image-based methods
to study neural circuitry that spans multiple brain regions, for quantifying brain tissue to classify cell types (neurons,
such as cortical-hippocampal[374] or dorsal-ventral.[114] Further- astrocytes, microglia, and endothelial cells) and analyze their
more, as described above, there are now published protocols associations in cellular networks have been developed,[390] the
for differentiating iPSCs into glutamatergic, midbrain, and use of label-free methods for real-time monitoring of cells is
dopaminergic neurons, inhibitory neurons, astrocytes, and oli- generally preferred. Using MEA-based impedance spectroscopy
godendrocytes. Through the use of bioengineering approaches, methods, researchers were able to monitor cellular pathology in
cells can be combined into 3D scaffolds or patterned networks real-time through the degradation of neurites in response to tau
to form circuit connections that can be used to examine func- hyperphosphorylation in slice cultures.[391] However, because
tional communication between neuronal subtypes and sup- MEAs generally offer a 2D measurement of a 3D tissue, recent
port cells.[375,376] Using cells isolated from healthy or patients efforts have sought to use optical methods such as multiphoton
afflicted with the disease, researchers can probe mechanisms microscopy for noncontact and noninvasive functional imaging
of circuit dysfunction at the scale of actual tissues in the of bioengineered neural tissues.[392] Wide-volume imaging
native 3D environment, and examine genetic or therapeutic demonstrated single cell resolution of neuronal activity via the

Adv. Funct. Mater. 2020, 1909146 1909146  (18 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

expression of a calcium sensor (GCaMP6s). By using single Keywords


and layered co-cultures of hippocampal and cortical neurons,
bioengineering, neurodegenerative and psychiatric diseases,
researchers were able to examine cell-type specific and inter-
neurodevelopment, organoids, stem cells
active neural network activity.[270] Optogenetic techniques are
simi­larly being optimized for noninvasive stimulation and Received: November 3, 2019
recording of electrophysiological activity. The further optimi- Revised: January 24, 2020
zation of technologies such as these for volumetric functional Published online:
imaging with high resolution will continue to drive neurosci-
ence discovery, particularly for high-throughput screening
applications.
[1] R. L. Buckner, F. M. Krienen, Trends Cognit. Sci. 2013, 17, 648.
[2] J. Defelipe, Front. Neuroanat. 2011, 5, 29.
[3] E. J. Nestler, S. E. Hyman, Nat. Neurosci. 2010, 13, 1161.
7. Conclusion [4] A. Pinson, T. Namba, W. B. Huttner, Front. Cell. Neurosci. 2019, 13, 305.
[5] M. Laubach, L. M. Amarante, K. Swanson, S. R. White, eNeuro
The emergence of 3D in vitro models of the human brain as 2018, 5, e.0315.
an alternative to traditional 2D culture systems offers a more [6] R. E. Andersen, D. A. Lim, Cell Tissue Res. 2018, 371, 55.
representative environment to observe and manipulate cel- [7] I. Bartha, J. di Iulio, J. C. Venter, A. Telenti, Nat. Rev. Genet. 2018,
lular behavior. To date, organoid technology or bioengineering 19, 51.
approaches have been the primary tools and techniques used [8] T. C. Burns, M. D. Li, S. Mehta, A. J. Awad, A. A. Morgan, Eur. J.
to prepare these 3D brain models. Though only in existence Pharmacol. 2015, 759, 101.
for a few years, organoids have proven their utility as a tool [9] T. F. Galatro, I. R. Holtman, A. M. Lerario, I. D. Vainchtein,
to model human development, function, disease, and evolu- N. Brouwer, P. R. Sola, M. M. Veras, T. F. Pereira, R. E. P. Leite,
T. Moller, P. D. Wes, M. C. Sogayar, J. D. Laman, W. den Dunnen,
tion, with early examples of therapeutic applications in neuro-
C. A. Pasqualucci, S. M. Oba-Shinjo, E. Boddeke, S. K. N. Marie,
logical and infectious disease and cancer biology. Despite this
B. J. L. Eggen, Nat. Neurosci. 2017, 20, 1162.
success, issues with reproducibility, heterogeneity, scalability, [10] D. Gosselin, D. Skola, N. G. Coufal, I. R. Holtman,
and necrosis continue to plague these cultures given that J. C. M. Schlachetzki, E. Sajti, B. N. Jaeger, C. O’Connor,
their formation and development are largely driven by self- C. Fitzpatrick, M. P. Pasillas, M. Pena, A. Adair, D. D. Gonda,
assembly. Bioengineering methodology, on the other hand, has M. L. Levy, R. M. Ransohoff, F. H. Gage, C. K. Glass, Science 2017,
been used to mimic microenvironments of the brain via cell- 356, eaal3222.
instructive hydrogels or scaffolding systems and/or connect [11] M. Hawrylycz, J. A. Miller, V. Menon, D. Feng, T. Dolbeare,
with other tissues through the use of chip-based microfluidic A. L. Guillozet-Bongaarts, A. G. Jegga, B. J. Aronow, C. K. Lee,
systems and bioreactors. While these approaches provide for a A. Bernard, M. F. Glasser, D. L. Dierker, J. Menche, A. Szafer,
F. Collman, P. Grange, K. A. Berman, S. Mihalas, Z. Yao,
greater level of control and reproducibility, they often lack the
L. Stewart, A. L. Barabasi, J. Schulkin, J. Phillips, L. Ng, C. Dang,
inherent complexity and physiological characteristics of the
D. R. Haynor, A. Jones, D. C. Van Essen, C. Koch, E. Lein, Nat.
native microenvironment. Given the state of the field, there is Neurosci. 2015, 18, 1832.
a clear opportunity to combine these approaches to generate [12] R. D. Hodge, T. E. Bakken, J. A. Miller, K. A. Smith, E. R. Barkan,
advanced 3D in vitro models with greater fidelity to the human L. T. Graybuck, J. L. Close, B. Long, N. Johansen, O. Penn, Z. Yao,
brain and enhanced utility in disease modeling, biomarker J. Eggermont, T. Hollt, B. P. Levi, S. I. Shehata, B. Aevermann,
identification, drug screening, and regenerative medicine. A. Beller, D. Bertagnolli, K. Brouner, T. Casper, C. Cobbs, R. Dalley,
Through the convergence of interdisciplinary fields including N. Dee, S. L. Ding, R. G. Ellenbogen, O. Fong, E. Garren, J. Goldy,
biology, neuroscience, biomedical engineering, and materials R. P. Gwinn, D. Hirschstein, et al., Nature 2019, 573, 61.
science, among others, we expect that future research will lead [13] G. La Manno, D. Gyllborg, S. Codeluppi, K. Nishimura, C. Salto,
A. Zeisel, L. E. Borm, S. R. W. Stott, E. M. Toledo, J. C. Villaescusa,
to better 3D models of the human brain that are suitable for
P. Lonnerberg, J. Ryge, R. A. Barker, E. Arenas, S. Linnarsson, Cell
high-throughput screening and compatible with technology
2016, 167, 566.
for real-time functional monitoring. Through these efforts, [14] T. Masuda, R. Sankowski, O. Staszewski, C. Bottcher, L. Amann,
researchers may finally be able to address the current lack of Sagar, C. Scheiwe, S. Nessler, P. Kunz, G. van Loo, V. A. Coenen,
treatment options and offer solutions for patients currently suf- P. C. Reinacher, A. Michel, U. Sure, R. Gold, D. Grun, J. Priller,
fering from neurological disease. C. Stadelmann, M. Prinz, Nature 2019, 566, 388.
[15] Y. Zhang, C. Pak, Y. Han, H. Ahlenius, Z. Zhang, S. Chanda,
S. Marro, C. Patzke, C. Acuna, J. Covy, W. Xu, N. Yang, T. Danko,
L. Chen, M. Wernig, T. C. Sudhof, Neuron 2013, 78, 785.
Acknowledgements [16] A. Bayes, M. O. Collins, M. D. Croning, L. N. van de Lagemaat,
M.L.L. and T.J.F.N. contributed equally to this work. The authors thank J. S. Choudhary, S. G. Grant, PLoS One 2012, 7, e46683.
the NIH (P41EB002520, R01NS092847, R01NS094218) for their support [17] K. Bozek, Y. Wei, Z. Yan, X. Liu, J. Xiong, M. Sugimoto, M. Tomita,
of this work. S. Paabo, C. C. Sherwood, P. R. Hof, J. J. Ely, Y. Li, D. Steinhauser,
L. Willmitzer, P. Giavalisco, P. Khaitovich, Neuron 2015, 85, 695.
[18] R. M. Ransohoff, J. Exp. Med. 2018, 215, 2955.
[19] C. D. Luft, E. Pereda, M. J. Banissy, J. Bhattacharya, Front. Syst.
Conflict of Interest Neurosci. 2014, 8, 132.
The authors declare no conflict of interest. [20] C. M. Michel, T. Koenig, NeuroImage 2018, 180, 577.

Adv. Funct. Mater. 2020, 1909146 1909146  (19 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[21] B. E. Youngerman, F. A. Khan, G. M. McKhann, Neuropsychiatr. [47] C. A. Tagge, A. M. Fisher, O. V. Minaeva, A. Gaudreau-Bal
Dis. Treat. 2019, 15, 1701. derrama, J. A. Moncaster, X. L. Zhang, M. W. Wojnarowicz,
[22] T. Kaufmann, D. van der Meer, N. T. Doan, E. Schwarz, N. Casey, H. Lu, O. N. Kokiko-Cochran, S. Saman, M. Ericsson,
M. J. Lund, I. Agartz, D. Alnaes, D. M. Barch, R. Baur-Streubel, K. D. Onos, R. Veksler, V. V. Senatorov Jr., A. Kondo, X. Z. Zhou,
A. Bertolino, F. Bettella, M. K. Beyer, E. Boen, S. Borgwardt, O. Miry, L. R. Vose, K. R. Gopaul, C. Upreti, C. J. Nowinski,
C. L. Brandt, J. Buitelaar, E. G. Celius, S. Cervenka, A. Conzelmann, R. C. Cantu, V. E. Alvarez, A. M. Hildebrandt, E. S. Franz, J. Konrad,
A. Cordova-Palomera, A. M. Dale, D. J. F. de Quervain, P. Carlo, J. A. Hamilton, N. Hua, Y. Tripodis, et al., Brain 2018, 141, 422.
S. Djurovic, E. S. Dorum, S. Eisenacher, T. Elvsashagen, T. Espeseth, [48] C. Lee, E. Y. Kang, M. J. Gandal, E. Eskin, D. H. Geschwind, Nat.
H. Fatouros-Bergman, L. Flyckt, et al., Nat. Neurosci. 2019, 22, 1617. Neurosci. 2019, 22, 1521.
[23] R. G. Rowe, G. Q. Daley, Nat. Rev. Genet. 2019, 20, 377. [49] M. S. Breen, A. Dobbyn, Q. Li, P. Roussos, G. E. Hoffman,
[24] L. Cong, F. A. Ran, D. Cox, S. Lin, R. Barretto, N. Habib, P. D. Hsu, E. Stahl, A. Chess, P. Sklar, J. B. Li, B. Devlin, J. D. Buxbaum, Nat.
X. Wu, W. Jiang, L. A. Marraffini, F. Zhang, Science 2013, 339, 819. Neurosci. 2019, 22, 1402.
[25] L. S. Qi, M. H. Larson, L. A. Gilbert, J. A. Doudna, J. S. Weissman, [50] V. V. Giridharan, P. Sayana, O. F. Pinjari, N. Ahmad, M. I.  da Rosa,
A. P. Arkin, W. A. Lim, Cell 2013, 152, 1173. J. Quevedo, T. Barichello, Mol. Psychiatry 2020, 25, 94.
[26] M. Jinek, K. Chylinski, I. Fonfara, M. Hauer, J. A. Doudna, [51] C. F. van Kesteren, H. Gremmels, L. D. de Witte, E. M. Hol,
E. Charpentier, Science 2012, 337, 816. A. R. Van Gool, P. G. Falkai, R. S. Kahn, I. E. Sommer, Transl.
[27] J. Engreitz, O. Abudayyeh, J. Gootenberg, F. Zhang, Cold Spring Psychiatry 2017, 7, e1075.
Harbor Perspect. Biol. 2019, 11, a035386. [52] R. E. McCullumsmith, J. H. Hammond, D. Shan,
[28] T. Gaj, C. A. Gersbach, C. F. Barbas 3rd, Trends Biotechnol. 2013, J. H. Meador-Woodruff, Neuropsychopharmacology 2014, 39, 65.
31, 397. [53] M. Freund, A. Taylor, C. Ng, A. R. Little, Handb. Clin. Neurol. 2018,
[29] M. Frega, M. Tedesco, P. Massobrio, M. Pesce, S. Martinoia, Sci. 150, 41.
Rep. 2014, 4, 5489. [54] D. B. Howieson, A. Dame, R. Camicioli, G. Sexton, H. Payami,
[30] H. Tekin, S. Simmons, B. Cummings, L. Gao, X. Adiconis, J. A. Kaye, J. Am. Geriatr. Soc. 1997, 45, 584.
C. C. Hession, A. Ghoshal, D. Dionne, S. R. Choudhury, [55] C. Humpel, Neuroscience 2015, 305, 86.
V. Yesilyurt, N. E. Sanjana, X. Shi, C. Lu, M. Heidenreich, J. Q. Pan, [56] C. L. Croft, M. A. Wade, K. Kurbatskaya, P. Mastrandreas,
J. Z. Levin, F. Zhang, Nat. Biomed. Eng. 2018, 2, 540. M. M. Hughes, E. C. Phillips, A. M. Pooler, M. S. Perkinton,
[31] J. Jo, Y. Xiao, A. X. Sun, E. Cukuroglu, H. D. Tran, J. Goke, D. P. Hanger, W. Noble, Cell Death Dis. 2017, 8, e2671.
Z. Y. Tan, T. Y. Saw, C. P. Tan, H. Lokman, Y. Lee, D. Kim, H. S. Ko, [57] Z. Yu, O. Graudejus, C. Tsay, S. P. Lacour, S. Wagner,
S. O. Kim, J. H. Park, N. J. Cho, T. M. Hyde, J. E. Kleinman, B. Morrison 3rd, J. Neurotrauma 2009, 26, 1135.
J. H. Shin, D. R. Weinberger, E. K. Tan, H. S. Je, H. H. Ng, Cell [58] Q. Li, X. Han, J. Wang, Mol. Neurobiol. 2016, 53, 4226.
Stem Cell 2016, 19, 248. [59] S. Ooba, H. Hasuo, M. Shigemori, S. Yamashita, T. Akasu, Kurume
[32] Y. H. Kim, S. H. Choi, C. D’Avanzo, M. Hebisch, C. Sliwinski, Med. J. 2009, 56, 49.
E. Bylykbashi, K. J. Washicosky, J. B. Klee, O. Brustle, R. E. Tanzi, [60] K. Harris, S. P. Armstrong, R. Campos-Pires, L. Kiru, N. P. Franks,
D. Y. Kim, Nat. Protoc. 2015, 10, 985. R. Dickinson, Anesthesiology 2013, 119, 1137.
[33] C. A. Trujillo, R. Gao, P. D. Negraes, J. Gu, J. Buchanan, S. Preissl, [61] J. Breder, C. F. Sabelhaus, T. Opitz, K. G. Reymann, U. H. Schroder,
A. Wang, W. Wu, G. G. Haddad, I. A. Chaim, A. Domissy, Neuropharmacology 2000, 39, 1779.
M. Vandenberghe, A. Devor, G. W. Yeo, B. Voytek, A. R. Muotri, [62] U. H. Schroder, J. Breder, C. F. Sabelhaus, K. G. Reymann, Neurop-
Cell Stem Cell 2019, 25, 558. harmacology 1999, 38, 319.
[34] T. H. Wen, D. K. Binder, I. M. Ethell, K. A. Razak, Front. Mol. Neu- [63] M. R. Lamprecht, B. Morrison 3rd, J. Neurotrauma 2015, 32, 1361.
rosci. 2018, 11, 270. [64] S. Shaefi, A. M. Mittel, J. A. Hyam, M. D. Boone, C. C. Chen,
[35] M. A. Lancaster, M. Renner, C. A. Martin, D. Wenzel, L. S. Bicknell, E. M. Kasper, Surg. Neurol. Int. 2016, 7, 103.
M. E. Hurles, T. Homfray, J. M. Penninger, A. P. Jackson, [65] M. A. Marques-Torrejon, E. Gangoso, S. M. Pollard, Dis. Models
J. A. Knoblich, Nature 2013, 501, 373. Mech. 2018, 11, dmm031435.
[36] P. Arlotta, S. P. Pasca, Curr. Opin. Neurobiol. 2019, 56, 194. [66] N. Priego, L. Zhu, C. Monteiro, M. Mulders, D. Wasilewski,
[37] T. Kadoshima, H. Sakaguchi, T. Nakano, M. Soen, S. Ando, W. Bindeman, L. Doglio, L. Martinez, E. Martinez-Saez,
M. Eiraku, Y. Sasai, Proc. Natl. Acad. Sci. USA 2013, 110, Y. C. S. Ramon, D. Megias, E. Hernandez-Encinas, C. Blanco-Aparicio,
20284. L. Martinez, E. Zarzuela, J. Munoz, C. Fustero-Torre, E. Pineiro-Yanez,
[38] G. D. Mahumane, P. Kumar, L. C.  du Toit, Y. E. Choonara, V. Pillay, A. Hernandez-Lain, L. Bertero, V. Poli, M. Sanchez-Martinez,
Biomater. Sci. 2018, 6, 2812. J. A. Menendez, R. Soffietti, J. Bosch-Barrera, M. Valiente, Nat. Med.
[39] A. Shafiee, A. Atala, Annu. Rev. Med. 2017, 68, 29. 2018, 24, 1024.
[40] M. A. Lancaster, N. S. Corsini, S. Wolfinger, E. H. Gustafson, [67] T. Eisemann, B. Costa, J. Strelau, M. Mittelbronn, P. Angel,
A. W. Phillips, T. R. Burkard, T. Otani, F. J. Livesey, J. A. Knoblich, H. Peterziel, BMC Cancer 2018, 18, 103.
Nat. Biotechnol. 2017, 35, 659. [68] J. Marksteiner, C. Humpel, Mol. Psychiatry 2008, 13, 939.
[41] S. Genon, A. Reid, R. Langner, K. Amunts, S. B. Eickhoff, Trends [69] C. R. Romero-Leguizamon, M. R. Elnagar, U. Kristiansen,
Cognit. Sci. 2018, 22, 350. K. A. Kohlmeier, Sci. Rep. 2019, 9, 1486.
[42] C. M. Schumann, C. W. Nordahl, Brain Res. 2011, 1380, 175. [70] N. Schwarz, U. B. S. Hedrich, H. Schwarz, A. H. P, N. Dammeier,
[43] J. Verheijen, K. Sleegers, Trends Genet. 2018, 34, 434. E. Auffenberg, F. Bedogni, J. B. Honegger, H. Lerche, T. V. Wuttke,
[44] J. L. Del-Aguila, Z. Li, U. Dube, K. A. Mihindukulasuriya, H. Koch, Sci. Rep. 2017, 7, 12249.
J. P. Budde, M. V. Fernandez, L. Ibanez, J. Bradley, F. Wang, [71] J. Rakotomamonjy, A. M. Ghoumari, Int. J. Mol. Sci. 2019, 20, 285.
K. Bergmann, R. Davenport, J. C. Morris, D. M. Holtzman, [72] S. Jang, H. Kim, H. J. Kim, S. K. Lee, E. W. Kim, K. Namkoong,
R. J. Perrin, B. A. Benitez, J. Dougherty, C. Cruchaga, O. Harari, E. Kim, Psychiatry Invest. 2018, 15, 205.
Alzheimer’s Res. Ther. 2019, 11, 71. [73] E. G. Z. Centeno, H. Cimarosti, A. Bithell, Mol. Neurodegener.
[45] D. Toulorge, A. H. Schapira, R. Hajj, J. Neurochem. 2016, 139, 27. 2018, 13, 27.
[46] A. Mallach, M. Weinert, J. Arthur, D. Gveric, T. S. Tierney, [74] L. Li, J. Chao, Y. Shi, Cell Tissue Res. 2018, 371, 143.
K. N. Alavian, FASEB J. 2019, 33, 6957. [75] J. Penney, W. T. Ralvenius, L. H. Tsai, Mol. Psychiatry 2020, 25, 148.

Adv. Funct. Mater. 2020, 1909146 1909146  (20 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[76] R. E. Marioni, S. E. Harris, Q. Zhang, A. F. McRae, E. S. Boyden, J. W. Lichtman, Z. M. Williams, S. A. McCarroll,


S. P. Hagenaars, W. D. Hill, G. Davies, C. W. Ritchie, C. R. Gale, P. Arlotta, Nature 2017, 545, 48.
J. M. Starr, A. M. Goate, D. J. Porteous, J. Yang, K. L. Evans, [95] J. G. Camp, F. Badsha, M. Florio, S. Kanton, T. Gerber,
I. J. Deary, N. R. Wray, P. M. Visscher, Transl. Psychiatry 2018, 8, 99. M. Wilsch-Brauninger, E. Lewitus, A. Sykes, W. Hevers,
[77] I. E. Jansen, J. E. Savage, K. Watanabe, J. Bryois, D. M. Williams, M. Lancaster, J. A. Knoblich, R. Lachmann, S. Paabo, W. B. Huttner,
S. Steinberg, J. Sealock, I. K. Karlsson, S. Hagg, L. Athanasiu, B. Treutlein, Proc. Natl. Acad. Sci. USA 2015, 112, 15672.
N. Voyle, P. Proitsi, A. Witoelar, S. Stringer, D. Aarsland, [96] C. Luo, M. A. Lancaster, R. Castanon, J. R. Nery, J. A. Knoblich,
I. S. Almdahl, F. Andersen, S. Bergh, F. Bettella, S. Bjornsson, J. R. Ecker, Cell Rep. 2016, 17, 3369.
A. Braekhus, G. Brathen, C. de Leeuw, R. S. Desikan, S. Djurovic, [97] L. Subramanian, M. Bershteyn, M. F. Paredes, A. R. Kriegstein,
L. Dumitrescu, T. Fladby, T. J. Hohman, P. V. Jonsson, S. J. Kiddle, Nat. Commun. 2017, 8, 14167.
et al., Nat. Genet. 2019, 51, 404. [98] A. M. Pasca, S. A. Sloan, L. E. Clarke, Y. Tian, C. D. Makinson,
[78] B. W. Kunkle, B. Grenier-Boley, R. Sims, J. C. Bis, V. Damotte, N. Huber, C. H. Kim, J. Y. Park, N. A. O’Rourke, K. D. Nguyen,
A. C. Naj, A. Boland, M. Vronskaya, S. J. van der Lee, A. Amlie-Wolf, S. J. Smith, J. R. Huguenard, D. H. Geschwind, B. A. Barres,
C. Bellenguez, A. Frizatti, V. Chouraki, E. R. Martin, K. Sleegers, S. P. Pasca, Nat. Methods 2015, 12, 671.
N. Badarinarayan, J. Jakobsdottir, K. L. Hamilton-Nelson, [99] T. Otani, M. C. Marchetto, F. H. Gage, B. D. Simons, F. J. Livesey,
S. Moreno-Grau, R. Olaso, R. Raybould, Y. Chen, A. B. Kuzma, Cell Stem Cell 2016, 18, 467.
M. Hiltunen, T. Morgan, S. Ahmad, B. N. Vardarajan, J. Epelbaum, [100] F. Mora-Bermudez, F. Badsha, S. Kanton, J. G. Camp, B. Vernot,
P. Hoffmann, M. Boada, et al., Nat. Genet. 2019, 51, 414. K. Kohler, B. Voigt, K. Okita, T. Maricic, Z. He, R. Lachmann,
[79] H. K. Lee, C. Velazquez Sanchez, M. Chen, P. J. Morin, J. M. Wells, S. Paabo, B. Treutlein, W. B. Huttner, eLife 2016, 5, e18683.
E. B. Hanlon, W. Xia, PLoS One 2016, 11, e0163072. [101] Y. Li, J. Muffat, A. Omer, I. Bosch, M. A. Lancaster, M. Sur,
[80] A. A. Sproul, S. Jacob, D. Pre, S. H. Kim, M. W. Nestor, L. Gehrke, J. A. Knoblich, R. Jaenisch, Cell Stem Cell 2017, 20, 385.
M. Navarro-Sobrino, I. Santa-Maria, M. Zimmer, S. Aubry, [102] A. Cardenas, A. Villalba, C. de Juan Romero, E. Pico, C. Kyrousi,
J. W. Steele, D. J. Kahler, A. Dranovsky, O. Arancio, J. F. Crary, A. C. Tzika, M. Tessier-Lavigne, L. Ma, M. Drukker, S. Cappello,
S. Gandy, S. A. Noggle, PLoS One 2014, 9, e84547. V. Borrell, Cell 2018, 174, 590.
[81] T. Yagi, D. Ito, Y. Okada, W. Akamatsu, Y. Nihei, T. Yoshizaki, [103] I. T. Fiddes, G. A. Lodewijk, M. Mooring, C. M. Bosworth,
S. Yamanaka, H. Okano, N. Suzuki, Hum. Mol. Genet. 2011, 20, A. D. Ewing, G. L. Mantalas, A. M. Novak, A. van den Bout,
4530. A. Bishara, J. L. Rosenkrantz, R. Lorig-Roach, A. R. Field,
[82] H. X. Deng, Y. Shi, Y. Yang, K. B. Ahmeti, N. Miller, C. Huang, M. Haeussler, L. Russo, A. Bhaduri, T. J. Nowakowski, A. A. Pollen,
L. Cheng, H. Zhai, S. Deng, K. Nuytemans, N. J. Corbett, M. J. Kim, M. L. Dougherty, X. Nuttle, M. C. Addor, S. Zwolinski, S. Katzman,
H. Deng, B. Tang, Z. Yang, Y. Xu, P. Chan, B. Huang, X. P. Gao, A. Kriegstein, E. E. Eichler, S. R. Salama, F. M. J. Jacobs,
Z. Song, Z. Liu, F. Fecto, N. Siddique, T. Foroud, J. Jankovic, D. Haussler, Cell 2018, 173, 1356.
B. Ghetti, D. A. Nicholson, D. Krainc, O. Melen, J. M. Vance, [104] A. C. O’Neill, C. Kyrousi, J. Klaus, R. J. Leventer, E. P. Kirk, A. Fry,
M. A. Pericak-Vance, Y. C. Ma, A. H. Rajput, T. Siddique, Nat. D. T. Pilz, T. Morgan, Z. A. Jenkins, M. Drukker, S. F. Berkovic,
Genet. 2016, 48, 733. I. E. Scheffer, R. Guerrini, D. M. Markie, M. Gotz, S. Cappello,
[83] R. Fernandez-Santiago, I. Carballo-Carbajal, G. Castellano, R. Torrent, S. P. Robertson, Cell Rep. 2018, 25, 2729.
Y. Richaud, A. Sanchez-Danes, R. Vilarrasa-Blasi, A. Sanchez-Pla, [105] A. Amiri, G. Coppola, S. Scuderi, F. Wu, T. Roychowdhury, F. Liu,
J. L. Mosquera, J. Soriano, J. Lopez-Barneo, J. M. Canals, J. Alberch, S. Pochareddy, Y. Shin, A. Safi, L. Song, Y. Zhu, A. M. M. Sousa,
A. Raya, M. Vila, A. Consiglio, J. I. Martin-Subero, M. Ezquerra, E. C. Psych, M. Gerstein, G. E. Crawford, N. Sestan, A. Abyzov,
E. Tolosa, EMBO Mol. Med. 2015, 7, 1529. F. M. Vaccarino, Science 2018, 362, eaat6720.
[84] S. R. Mehta, C. M. Tom, Y. Wang, C. Bresee, D. Rushton, [106] L. Wang, S. Hou, Y. G. Han, Nat. Neurosci. 2016, 19, 888.
P. P. Mathkar, J. Tang, V. B. Mattis, Cell Rep. 2018, 25, 1081. [107] B. A. Lindborg, J. H. Brekke, A. L. Vegoe, C. B. Ulrich,
[85] F. B. Russo, A. Brito, A. M. de Freitas, A. Castanha, B. C. de Freitas, K. T. Haider, S. Subramaniam, S. L. Venhuizen, C. R. Eide,
P. C. B. Beltrao-Braga, Neurobiol. Dis. 2019, 130, 104483. P. J. Orchard, W. Chen, Q. Wang, F. Pelaez, C. M. Scott, E. Kokkoli,
[86] S. H. Choi, Y. H. Kim, M. Hebisch, C. Sliwinski, S. Lee, S. A. Keirstead, J. R. Dutton, J. Tolar, T. D. O’Brien, Stem Cells
C. D’Avanzo, H. Chen, B. Hooli, C. Asselin, J. Muffat, J. B. Klee, Transl. Med. 2016, 5, 970.
C. Zhang, B. J. Wainger, M. Peitz, D. M. Kovacs, C. J. Woolf, [108] J. Mariani, G. Coppola, P. Zhang, A. Abyzov, L. Provini,
S. L. Wagner, R. E. Tanzi, D. Y. Kim, Nature 2014, 515, 274. L. Tomasini, M. Amenduni, A. Szekely, D. Palejev, M. Wilson,
[87] T. Nakano, S. Ando, N. Takata, M. Kawada, K. Muguruma, M. Gerstein, E. L. Grigorenko, K. Chawarska, K. A. Pelphrey,
K. Sekiguchi, K. Saito, S. Yonemura, M. Eiraku, Y. Sasai, Cell Stem J. R. Howe, F. M. Vaccarino, Cell 2015, 162, 375.
Cell 2012, 10, 771. [109] X. Qian, H. N. Nguyen, M. M. Song, C. Hadiono, S. C. Ogden,
[88] M. Eiraku, N. Takata, H. Ishibashi, M. Kawada, E. Sakakura, C. Hammack, B. Yao, G. R. Hamersky, F. Jacob, C. Zhong,
S. Okuda, K. Sekiguchi, T. Adachi, Y. Sasai, Nature 2011, 472, 51. K. J. Yoon, W. Jeang, L. Lin, Y. Li, J. Thakor, D. A. Berg,
[89] H. Suga, T. Kadoshima, M. Minaguchi, M. Ohgushi, M. Soen, C. Zhang, E. Kang, M. Chickering, D. Nauen, C. Y. Ho, Z. Wen,
T. Nakano, N. Takata, T. Wataya, K. Muguruma, H. Miyoshi, K. M. Christian, P. Y. Shi, B. J. Maher, H. Wu, P. Jin, H. Tang,
S. Yonemura, Y. Oiso, Y. Sasai, Nature 2011, 480, 57. H. Song, G. L. Ming, Cell 2016, 165, 1238.
[90] T. Sato, H. Clevers, Science 2013, 340, 1190. [110] A. S. Monzel, L. M. Smits, K. Hemmer, S. Hachi, E. L. Moreno,
[91] T. Sato, R. G. Vries, H. J. Snippert, M. van de Wetering, N. Barker, T. van Wuellen, J. Jarazo, J. Walter, I. Bruggemann, I. Boussaad,
D. E. Stange, J. H. van Es, A. Abo, P. Kujala, P. J. Peters, E. Berger, R. M. T. Fleming, S. Bolognin, J. C. Schwamborn, Stem
H. Clevers, Nature 2009, 459, 262. Cell Rep. 2017, 8, 1144.
[92] J. R. Spence, C. N. Mayhew, S. A. Rankin, M. F. Kuhar, [111] J. A. Bagley, D. Reumann, S. Bian, J. Levi-Strauss, J. A. Knoblich,
J. E. Vallance, K. Tolle, E. E. Hoskins, V. V. Kalinichenko, S. I. Wells, Nat. Methods 2017, 14, 743.
A. M. Zorn, N. F. Shroyer, J. M. Wells, Nature 2011, 470, 105. [112] X. Qian, F. Jacob, M. M. Song, H. N. Nguyen, H. Song, G. L. Ming,
[93] M. A. Lancaster, J. A. Knoblich, Nat. Protoc. 2014, 9, 2329. Nat. Protoc. 2018, 13, 565.
[94] G. Quadrato, T. Nguyen, E. Z. Macosko, J. L. Sherwood, S. Min [113] Y. Yan, L. Song, J. Madinya, T. Ma, Y. Li, Tissue Eng., Part A 2018,
Yang, D. R. Berger, N. Maria, J. Scholvin, M. Goldman, J. P. Kinney, 24, 418.

Adv. Funct. Mater. 2020, 1909146 1909146  (21 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[114] F. Birey, J. Andersen, C. D. Makinson, S. Islam, W. Wei, [133] P. Conforti, D. Besusso, V. D. Bocchi, A. Faedo, E. Cesana,
N. Huber, H. C. Fan, K. R. C. Metzler, G. Panagiotakos, N. Thom, G. Rossetti, V. Ranzani, C. N. Svendsen, L. M. Thompson,
N. A. O’Rourke, L. M. Steinmetz, J. A. Bernstein, J. Hallmayer, M. Toselli, G. Biella, M. Pagani, E. Cattaneo, Proc. Natl. Acad. Sci.
J. R. Huguenard, S. P. Pasca, Nature 2017, 545, 54. USA 2018, 115, E762.
[115] S. J. Yoon, L. S. Elahi, A. M. Pasca, R. M. Marton, A. Gordon, [134] P. Wang, R. Mokhtari, E. Pedrosa, M. Kirschenbaum, C. Bayrak,
O. Revah, Y. Miura, E. M. Walczak, G. M. Holdgate, H. C. Fan, D. Zheng, H. M. Lachman, Mol. Autism 2017, 8, 11.
J. R. Huguenard, D. H. Geschwind, S. P. Pasca, Nat. Methods [135] M. T. Siu, D. T. Butcher, A. L. Turinsky, C. Cytrynbaum,
2019, 16, 75. D. J. Stavropoulos, S. Walker, O. Caluseriu, M. Carter, Y. Lou,
[116] S. A. Sloan, J. Andersen, A. M. Pasca, F. Birey, S. P. Pasca, Nat. R. Nicolson, S. Georgiades, P. Szatmari, E. Anagnostou,
Protoc. 2018, 13, 2062. S. W. Scherer, S. Choufani, M. Brudno, R. Weksberg, Clin. Epigenet.
[117] V. Tieng, L. Stoppini, S. Villy, M. Fathi, M. Dubois-Dauphin, 2019, 11, 103.
K. H. Krause, Stem Cells Dev. 2014, 23, 1535. [136] K. J. Yoon, F. R. Ringeling, C. Vissers, F. Jacob, M. Pokrass,
[118] S. Velasco, A. J. Kedaigle, S. K. Simmons, A. Nash, M. Rocha, D. Jimenez-Cyrus, Y. Su, N. S. Kim, Y. Zhu, L. Zheng, S. Kim,
G. Quadrato, B. Paulsen, L. Nguyen, X. Adiconis, A. Regev, X. Wang, L. C. Dore, P. Jin, S. Regot, X. Zhuang, S. Canzar, C. He,
J. Z. Levin, P. Arlotta, Nature 2019, 570, 523. G. L. Ming, H. Song, Cell 2017, 171, 877.
[119] M. Schukking, H. C. Miranda, C. A. Trujillo, P. D. Negraes, [137] M. L. Allende, E. K. Cook, B. C. Larman, A. Nugent, J. M. Brady,
A. R. Muotri, Stem Cells Dev. 2018, 27, 1549. D. Golebiowski, M. Sena-Esteves, C. J. Tifft, R. L. Proia, J. Lipid Res.
[120] V. S. Sohal, J. L. R. Rubenstein, Mol. Psychiatry 2019, 24, 1248. 2018, 59, 550.
[121] F. Ariani, G. Hayek, D. Rondinella, R. Artuso, M. A. Mencarelli, [138] M. Jin, O. Pomp, T. Shinoda, S. Toba, T. Torisawa, K. Furuta,
A. Spanhol-Rosseto, M. Pollazzon, S. Buoni, O. Spiga, S. Ricciardi, K. Oiwa, T. Yasunaga, D. Kitagawa, S. Matsumura, T. Miyata,
I. Meloni, I. Longo, F. Mari, V. Broccoli, M. Zappella, A. Renieri, T. T. Tan, B. Reversade, S. Hirotsune, Sci. Rep. 2017, 7, 39902.
Am. J. Hum. Genet. 2008, 83, 89. [139] V. Iefremova, G. Manikakis, O. Krefft, A. Jabali, K. Weynans,
[122] N. Bahi-Buisson, J. Nectoux, B. Girard, H. Van Esch, T. De Ravel, R. Wilkens, F. Marsoner, B. Brandl, F. J. Muller, P. Koch,
N. Boddaert, P. Plouin, M. Rio, Y. Fichou, J. Chelly, T. Bienvenu, J. Ladewig, Cell Rep. 2017, 19, 50.
Neurogenetics 2010, 11, 241. [140] M. Bershteyn, T. J. Nowakowski, A. A. Pollen, E. Di Lullo, A. Nene,
[123] S. A. Shoichet, S. A. Kunde, P. Viertel, C. Schell-Apacik, A. Wynshaw-Boris, A. R. Kriegstein, Cell Stem Cell 2017, 20, 435.
H. von Voss, N. Tommerup, H. H. Ropers, V. M. Kalscheuer, Hum. [141] R. Li, L. Sun, A. Fang, P. Li, Q. Wu, X. Wang, Protein Cell 2017, 8,
Genet. 2005, 117, 536. 823.
[124] N. Mellios, D. A. Feldman, S. D. Sheridan, J. P. K. Ip, S. Kwok, [142] T. Matsui, V. Nieto-Estevez, S. Kyrychenko, J. W. Schneider,
S. K. Amoah, B. Rosen, B. A. Rodriguez, B. Crawford, J. Hsieh, Development 2017, 144, 1025.
R. Swaminathan, S. Chou, Y. Li, M. Ziats, C. Ernst, R. Jaenisch, [143] Y. Xiang, Y. Tanaka, B. Patterson, Y. J. Kang, G. Govindaiah,
S. J. Haggarty, M. Sur, Mol. Psychiatry 2018, 23, 1051. N. Roselaar, B. Cakir, K. Y. Kim, A. P. Lombroso, S. M. Hwang,
[125] A. A. Pollen, A. Bhaduri, M. G. Andrews, T. J. Nowakowski, M. Zhong, E. G. Stanley, A. G. Elefanty, J. R. Naegele, S. H. Lee,
O. S. Meyerson, M. A. Mostajo-Radji, E. Di Lullo, B. Alvarado, S. M. Weissman, I. H. Park, Cell Stem Cell 2017, 21, 383.
M. Bedolli, M. L. Dougherty, I. T. Fiddes, Z. N. Kronenberg, [144] R. Tejero, Y. Huang, I. Katsyv, M. Kluge, J. Y. Lin, J. Tome-Garcia,
J. Shuga, A. A. Leyrat, J. A. West, M. Bershteyn, C. B. Lowe, N. Daviaud, Y. Wang, B. Zhang, N. M. Tsankova, C. C. Friedel,
B. J. Pavlovic, S. R. Salama, D. Haussler, E. E. Eichler, H. Zou, R. H. Friedel, EBioMedicine 2019, 42, 252.
A. R. Kriegstein, Cell 2019, 176, 743. [145] S. A. Sloan, S. Darmanis, N. Huber, T. A. Khan, F. Birey, C. Caneda,
[126] A. R. Field, F. M. J. Jacobs, I. T. Fiddes, A. P. R. Phillips, R. Reimer, S. R. Quake, B. A. Barres, S. P. Pasca, Neuron 2017, 95, 779.
A. M. Reyes-Ortiz, E. LaMontagne, L. Whitehead, V. Meng, [146] F. Ye, E. Kang, C. Yu, X. Qian, F. Jacob, C. Yu, M. Mao,
J. L. Rosenkrantz, M. Olsen, M. Hauessler, S. Katzman, S. R. Salama, R. Y. C. Poon, J. Kim, H. Song, G. L. Ming, M. Zhang, Neuron
D. Haussler, Stem Cell Rep. 2019, 12, 245. 2017, 96, 1041.
[127] X. Yang, B. Xu, B. Mulvey, M. Evans, S. Jordan, Y. D. Wang, [147] P. Srikanth, V. N. Lagomarsino, C. R. Muratore, S. C. Ryu, A. He,
V. Pagala, J. Peng, Y. Fan, A. Patel, J. C. Peng, Nat. Neurosci. 2019, W. M. Taylor, C. Zhou, M. Arellano, T. L. Young-Pearse, Transl.
22, 362. Psychiatry 2018, 8, 77.
[128] F. Perez-Branguli, I. Y. Buchsbaum, T. Pozner, M. Regensburger, [148] W. K. Raja, A. E. Mungenast, Y. T. Lin, T. Ko, F. Abdurrob, J. Seo,
W. Fan, A. Schray, T. Borstler, H. Mishra, D. Graf, Z. Kohl, L. H. Tsai, PLoS One 2016, 11, e0161969.
J. Winkler, B. Berninger, S. Cappello, B. Winner, Hum. Mol. Genet. [149] C. Gonzalez, E. Armijo, J. Bravo-Alegria, A. Becerra-Calixto,
2019, 28, 961. C. E. Mays, C. Soto, Mol. Psychiatry 2018, 23, 2363.
[129] J. Klaus, S. Kanton, C. Kyrousi, A. C. Ayo-Martin, R. Di Giaimo, [150] S. Pavoni, R. Jarray, F. Nassor, A. C. Guyot, S. Cottin, J. Rontard,
S. Riesenberg, A. C. O’Neill, J. G. Camp, C. Tocco, M. Santel, J. Mikol, A. Mabondzo, J. P. Deslys, F. Yates, PLoS One 2018, 13,
E. Rusha, M. Drukker, M. Schroeder, M. Gotz, S. P. Robertson, e0209150.
B. Treutlein, S. Cappello, Nat. Med. 2019, 25, 561. [151] Y. Yan, L. Song, J. Bejoy, J. Zhao, T. Kanekiyo, G. Bu, Y. Zhou, Y. Li,
[130] C. A. Thomas, L. Tejwani, C. A. Trujillo, P. D. Negraes, R. H. Herai, Tissue Eng., Part A 2018, 24, 1125.
P. Mesci, A. Macia, Y. J. Crow, A. R. Muotri, Cell Stem Cell 2017, [152] R. Xu, A. T. Brawner, S. Li, J. J. Liu, H. Kim, H. Xue, Z. P. Pang,
21, 319. W. Y. Kim, R. P. Hart, Y. Liu, P. Jiang, Cell Stem Cell 2019, 24, 908.
[131] E. K. Stachowiak, C. A. Benson, S. T. Narla, A. Dimitri, L. E. B. Chuye, [153] Y. T. Lin, J. Seo, F. Gao, H. M. Feldman, H. L. Wen, J. Penney,
S. Dhiman, K. Harikrishnan, S. Elahi, D. Freedman, K. J. Brennand, H. P. Cam, E. Gjoneska, W. K. Raja, J. Cheng, R. Rueda, O. Kritskiy,
P. Sarder, M. K. Stachowiak, Transl. Psychiatry 2017, 7, 6. F. Abdurrob, Z. Peng, B. Milo, C. J. Yu, S. Elmsaouri, D. Dey, T. Ko,
[132] M. E. Coulter, C. M. Dorobantu, G. A. Lodewijk, F. Delalande, B. A. Yankner, L. H. Tsai, Neuron 2018, 98, 1141.
S. Cianferani, V. S. Ganesh, R. S. Smith, E. T. Lim, C. S. Xu, [154] K. Meyer, H. M. Feldman, T. Lu, D. Drake, E. T. Lim, K. H. Ling,
S. Pang, E. T. Wong, H. G. W. Lidov, M. L. Calicchio, E. Yang, N. A. Bishop, Y. Pan, J. Seo, Y. T. Lin, S. C. Su, G. M. Church,
D. M. Gonzalez, T. M. Schlaeger, G. H. Mochida, H. Hess, L. H. Tsai, B. A. Yankner, Cell Rep. 2019, 26, 1112.
W. A. Lee, M. K. Lehtinen, T. Kirchhausen, D. Haussler, [155] E. M. Boisvert, R. E. Means, M. Michaud, J. A. Madri, S. G. Katz,
F. M. J. Jacobs, R. Gaudin, C. A. Walsh, Cell Rep. 2018, 24, 973. Cell Death Dis. 2019, 10, 325.

Adv. Funct. Mater. 2020, 1909146 1909146  (22 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[156] A. M. Pasca, J. Y. Park, H. W. Shin, Q. Qi, O. Revah, R. Krasnoff, D. Gong, X. Dai, R. Damoiseaux, R. Aliyari, S. Liebscher,
R. O’Hara, A. J. Willsey, T. D. Palmer, S. P. Pasca, Nat. Med. 2019, K. Schenke-Layland, C. Caneda, E. J. Huang, Y. Zhang, G. Cheng,
25, 784. D. H. Geschwind, P. Golshani, R. Sun, B. G. Novitch, Cell Rep.
[157] R. C. Sartore, S. C. Cardoso, Y. V. Lages, J. M. Paraguassu, 2017, 21, 517.
M. P. Stelling, R. F. Madeiro da Costa, M. Z. Guimaraes, [176] M. Xu, E. M. Lee, Z. Wen, Y. Cheng, W. K. Huang, X. Qian,
C. A. Perez, S. K. Rehen, PeerJ 2017, 5, e2927. J. Tcw, J. Kouznetsova, S. C. Ogden, C. Hammack, F. Jacob,
[158] F. Liu, J. Huang, Z. Liu, Neuroscience 2019, 404, 530. H. N. Nguyen, M. Itkin, C. Hanna, P. Shinn, C. Allen,
[159] C. T. Lee, J. Chen, A. A. Kindberg, R. M. Bendriem, C. E. Spivak, S. G. Michael, A. Simeonov, W. Huang, K. M. Christian, A. Goate,
M. P. Williams, C. T. Richie, A. Handreck, B. S. Mallon, K. J. Brennand, R. Huang, M. Xia, G. L. Ming, W. Zheng, H. Song,
C. R. Lupica, D. T. Lin, B. K. Harvey, D. C. Mash, W. J. Freed, H. Tang, Nat. Med. 2016, 22, 1101.
Neuropsychopharmacology 2017, 42, 774. [177] C. Q. Sacramento, G. R. de Melo, C. S. de Freitas, N. Rocha,
[160] Y. Wang, L. Wang, Y. Zhu, J. Qin, Lab Chip 2018, 18, 851. L. V. Hoelz, M. Miranda, N. Fintelman-Rodrigues, A. Marttorelli,
[161] Y. Zhu, L. Wang, F. Yin, Y. Yu, Y. Wang, M. J. Shepard, Z. Zhuang, A. C. Ferreira, G. Barbosa-Lima, J. L. Abrantes, Y. R. Vieira,
J. Qin, Integr. Biol. 2017, 9, 968. M. M. Bastos, E. de Mello Volotao, E. P. Nunes, D. A. Tschoeke,
[162] J. Huang, F. Liu, H. Tang, H. Wu, L. Li, R. Wu, J. Zhao, Y. Wu, L. Leomil, E. C. Loiola, P. Trindade, S. K. Rehen, F. A. Bozza,
Z. Liu, J. Chen, Front. Neurol. 2017, 8, 626. P. T. Bozza, N. Boechat, F. L. Thompson, A. M. de Filippis,
[163] V. Dakic, J. Minardi Nascimento, R. Costa Sartore, R. M. Maciel, K. Bruning, T. M. Souza, Sci. Rep. 2017, 7, 40920.
D. B. de Araujo, S. Ribeiro, D. Martins-de-Souza, S. K. Rehen, Sci. [178] C. Li, Y. Q. Deng, S. Wang, F. Ma, R. Aliyari, X. Y. Huang,
Rep. 2017, 7, 12863. N. N. Zhang, M. Watanabe, H. L. Dong, P. Liu, X. F. Li, Q. Ye,
[164] B. R. Groveman, S. T. Foliaki, C. D. Orru, G. Zanusso, J. A. Carroll, M. Tian, S. Hong, J. Fan, H. Zhao, L. Li, N. Vishlaghi, J. E. Buth,
B. Race, C. L. Haigh, Acta Neuropathol. Commun. 2019, 7, 90. C. Au, Y. Liu, N. Lu, P. Du, F. X. Qin, B. Zhang, D. Gong, X. Dai,
[165] L. D’Aiuto, D. C. Bloom, J. N. Naciri, A. Smith, T. G. Edwards, R. Sun, B. G. Novitch, Z. Xu, C. F. Qin, G. Cheng, Immunity 2017,
L. McClain, J. A. Callio, M. Jessup, J. Wood, K. Chowdari, 46, 446.
M. Demers, E. E. Abrahamson, M. D. Ikonomovic, L. Viggiano, [179] T. Zhou, L. Tan, G. Y. Cederquist, Y. Fan, B. J. Hartley, S. Mukherjee,
R. De Zio, S. Watkins, P. R. Kinchington, V. L. Nimgaonkar, J. Virol. M. Tomishima, K. J. Brennand, Q. Zhang, R. E. Schwartz, T. Evans,
2019, 93, e00111. L. Studer, S. Chen, Cell Stem Cell 2017, 21, 274.
[166] J. Muffat, Y. Li, A. Omer, A. Durbin, I. Bosch, G. Bakiasi, [180] Y. P. Xu, Y. Qiu, B. Zhang, G. Chen, Q. Chen, M. Wang, F. Mo,
E. Richards, A. Meyer, L. Gehrke, R. Jaenisch, Proc. Natl. Acad. Sci. J. Xu, J. Wu, R. R. Zhang, M. L. Cheng, N. N. Zhang, B. Lyu,
USA 2018, 115, 7117. W. L. Zhu, M. H. Wu, Q. Ye, D. Zhang, J. H. Man, X. F. Li, J. Cui,
[167] B. Zhang, Y. He, Y. Xu, F. Mo, T. Mi, Q. S. Shen, C. Li, Y. Li, J. Liu, Z. Xu, B. Hu, X. Zhou, C. F. Qin, Cell Res. 2019, 29, 265.
Y. Wu, G. Chen, W. Zhu, C. Qin, B. Hu, G. Zhou, Cell Death Dis. [181] H. Sakaguchi, Y. Ozaki, T. Ashida, T. Matsubara, N. Oishi,
2018, 9, 719. S. Kihara, J. Takahashi, Stem Cell Rep. 2019, 13, 458.
[168] P. P. Garcez, E. C. Loiola, R. Madeiro da Costa, L. M. Higa, [182] D. Poli, C. Magliaro, A. Ahluwalia, Front. Neurosci. 2019, 13, 162.
P. Trindade, R. Delvecchio, J. M. Nascimento, R. Brindeiro, [183] Y. Zhao, P. Liu, N. Zhang, J. Chen, L. D. Landegger, L. Wu, F. Zhao,
A. Tanuri, S. K. Rehen, Science 2016, 352, 816. Y. Zhao, Y. Zhang, J. Zhang, T. Fujita, A. Stemmer-Rachamimov,
[169] K. J. Yoon, G. Song, X. Qian, J. Pan, D. Xu, H. S. Rho, N. S. Kim, G. B. Ferraro, H. Liu, A. Muzikansky, S. R. Plotkin, K. M. Stankovic,
C. Habela, L. Zheng, F. Jacob, F. Zhang, E. M. Lee, W. K. Huang, R. K. Jain, L. Xu, Proc. Natl. Acad. Sci. USA 2018, 115, E2077.
F. R. Ringeling, C. Vissers, C. Li, L. Yuan, K. Kang, S. Kim, J. Yeo, [184] S. Bian, M. Repic, Z. Guo, A. Kavirayani, T. Burkard, J. A. Bagley,
Y. Cheng, S. Liu, Z. Wen, C. F. Qin, Q. Wu, K. M. Christian, C. Krauditsch, J. A. Knoblich, Nat. Methods 2018, 15, 631.
H. Tang, P. Jin, Z. Xu, J. Qian, et al., Cell Stem Cell 2017, 21, 349. [185] B. Decaesteker, G. Denecker, C. Van Neste, E. M. Dolman,
[170] F. R. Cugola, I. R. Fernandes, F. B. Russo, B. C. Freitas, J. L. Dias, W. Van Loocke, M. Gartlgruber, C. Nunes, F. De Vloed, P. Depuydt,
K. P. Guimaraes, C. Benazzato, N. Almeida, G. C. Pignatari, K. Verboom, D. Rombaut, S. Loontiens, J. De Wyn, W. M. Kholosy,
S. Romero, C. M. Polonio, I. Cunha, C. L. Freitas, W. N. Brandao, B. Koopmans, A. H. W. Essing, C. Herrmann, D. Dreidax, K. Durinck,
C. Rossato, D. G. Andrade, P. Faria Dde, A. T. Garcez, D. Deforce, F. Van Nieuwerburgh, A. Henssen, R. Versteeg, V. Boeva,
C. A. Buchpigel, C. T. Braconi, E. Mendes, A. A. Sall, P. M. Zanotto, G. Schleiermacher, J. van Nes, P. Mestdagh, S. Vanhauwaert,
J. P. Peron, A. R. Muotri, P. C. Beltrao-Braga, Nature 2016, 534, 267. J. H. Schulte, F. Westermann, et al., Nat. Commun. 2018, 9, 4866.
[171] E. Gabriel, A. Ramani, U. Karow, M. Gottardo, K. Natarajan, [186] S. Bergmann, S. E. Lawler, Y. Qu, C. M. Fadzen, J. M. Wolfe,
L. M. Gooi, G. Goranci-Buzhala, O. Krut, F. Peters, M. Nikolic, M. S. Regan, B. L. Pentelute, N. Y. R. Agar, C. F. Cho, Nat. Protoc.
S. Kuivanen, E. Korhonen, T. Smura, O. Vapalahti, A. Papantonis, 2018, 13, 2827.
J. Schmidt-Chanasit, M. Riparbelli, G. Callaini, M. Kronke, [187] Y. Xiang, T. Yoshiaki, B. Patterson, B. Cakir, K. Y. Kim, Y. S. Cho,
O. Utermohlen, J. Gopalakrishnan, Cell Stem Cell 2017, 20, 397. I. H. Park, Curr. Protoc. Stem Cell Biol. 2018, 47, e61.
[172] Y. X. Setoh, A. A. Amarilla, N. Y. G. Peng, R. E. Griffiths, J. Carrera, [188] R. M. Marton, Y. Miura, S. A. Sloan, Q. Li, O. Revah, R. J. Levy,
M. E. Freney, E. Nakayama, S. Ogawa, D. Watterson, N. Modhiran, J. R. Huguenard, S. P. Pasca, Nat. Neurosci. 2019, 22, 484.
F. E. Nanyonga, F. J. Torres, A. Slonchak, P. Periasamy, N. A. Prow, [189] H. Kim, R. Xu, R. Padmashri, A. Dunaevsky, Y. Liu, C. F. Dreyfus,
B. Tang, J. Harrison, J. Hobson-Peters, T. Cuddihy, J. Cooper-White, P. Jiang, Stem Cell Rep. 2019, 12, 890.
R. A. Hall, P. R. Young, J. M. Mackenzie, E. Wolvetang, J. D. Bloom, [190] B. S. Carter, G. S. Bova, T. H. Beaty, G. D. Steinberg, B. Childs,
A. Suhrbier, A. A. Khromykh, Nat. Microbiol. 2019, 4, 876. W. B. Isaacs, P. C. Walsh, J. Urol. 1993, 150, 797.
[173] S. Janssens, M. Schotsaert, R. Karnik, V. Balasubramaniam, [191] E. M. Abud, R. N. Ramirez, E. S. Martinez, L. M. Healy,
M. Dejosez, A. Meissner, A. Garcia-Sastre, T. P. Zwaka, mSystems C. H. H. Nguyen, S. A. Newman, A. V. Yeromin, V. M. Scarfone,
2018, 3, e00219. S. E. Marsh, C. Fimbres, C. A. Caraway, G. M. Fote, A. M. Madany,
[174] M. F. Wells, M. R. Salick, O. Wiskow, D. J. Ho, K. A. Worringer, A. Agrawal, R. Kayed, K. H. Gylys, M. D. Cahalan, B. J. Cummings,
R. J. Ihry, S. Kommineni, B. Bilican, J. R. Klim, E. J. Hill, L. T. Kane, J. P. Antel, A. Mortazavi, M. J. Carson, W. W. Poon, M. Blurton-Jones,
C. Ye, A. Kaykas, K. Eggan, Cell Stem Cell 2016, 19, 703. Neuron 2017, 94, 278.
[175] M. Watanabe, J. E. Buth, N. Vishlaghi, L. de la Torre-Ubieta, [192] P. R. Ormel, R. Vieira de Sa, E. J. van Bodegraven, H. Karst,
J. Taxidis, B. S. Khakh, G. Coppola, C. A. Pearson, K. Yamauchi, O. Harschnitz, M. A. M. Sneeboer, L. E. Johansen, R. E. van Dijk,

Adv. Funct. Mater. 2020, 1909146 1909146  (23 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

N. Scheefhals, A. Berdenis van Berlekom, E. Ribes Martinez, [216] Y. Shin, S. H. Choi, E. Kim, E. Bylykbashi, J. A. Kim, S. Chung,
S. Kling, H. D. MacGillavry, L. H. van den Berg, R. S. Kahn, D. Y. Kim, R. D. Kamm, R. E. Tanzi, Adv. Sci. 2019, 6, 1900962.
E. M. Hol, L. D. de Witte, R. J. Pasterkamp, Nat. Commun. 2018, [217] K. M. Fabre, L. Delsing, R. Hicks, N. Colclough, D. C. Crowther,
9, 4167. L. Ewart, Adv. Drug Delivery Rev. 2019, 140, 129.
[193] Y. Xiang, Y. Tanaka, B. Cakir, B. Patterson, K. Y. Kim, P. Sun, [218] M. Karimi, S. Bahrami, H. Mirshekari, S. M. M. Basri, A. B. Nik,
Y. J. Kang, M. Zhong, X. Liu, P. Patra, S. H. Lee, S. M. Weissman, A. R. Aref, M. Akbari, M. R. Hamblin, Lab Chip 2016, 16, 2551.
I. H. Park, Cell Stem Cell 2019, 24, 487. [219] J. Y. Park, S. K. Kim, D. H. Woo, E. J. Lee, J. H. Kim, S. H. Lee,
[194] A. A. Mansour, J. T. Goncalves, C. W. Bloyd, H. Li, S. Fernandes, Stem Cells 2009, 27, 2646.
D. Quang, S. Johnston, S. L. Parylak, X. Jin, F. H. Gage, Nat. [220] Y. Wang, L. Wang, Y. Guo, Y. Zhu, J. Qin, RSC Adv. 2018, 8, 1677.
Biotechnol. 2018, 36, 432. [221] E. Karzbrun, A. Kshirsagar, S. R. Cohen, J. H. Hanna, O. Reiner,
[195] N. Daviaud, R. H. Friedel, H. Zou, eNeuro 2018, 5, Nat. Phys. 2018, 14, 515.
ENEURO.0219-18.2018. [222] P. Zhuang, A. X. Sun, J. An, C. K. Chua, S. Y. Chew, Biomaterials
[196] D. Huh, H. Fujioka, Y. C. Tung, N. Futai, R. Paine 3rd, 2018, 154, 113.
J. B. Grotberg, S. Takayama, Proc. Natl. Acad. Sci. USA 2007, 104, [223] D. Simao, M. M. Silva, A. P. Terrasso, F. Arez, M. F. Q. Sousa,
18886. N. Z. Mehrjardi, T. Saric, P. Gomes-Alves, N. Raimundo,
[197] D. Huh, B. D. Matthews, A. Mammoto, M. Montoya-Zavala, P. M. Alves, C. Brito, Stem Cell Rep. 2018, 11, 552.
H. Y. Hsin, D. E. Ingber, Science 2010, 328, 1662. [224] M. E. Boutin, L. L. Kramer, L. L. Livi, T. Brown, C. Moore,
[198] H. Kimura, T. Yamamoto, H. Sakai, Y. Sakai, T. Fujii, Lab Chip D. Hoffman-Kim, J. Neurosci. Methods 2018, 299, 55.
2008, 8, 741. [225] G. Nzou, R. T. Wicks, E. E. Wicks, S. A. Seale, C. H. Sane, A. Chen,
[199] S. R. Khetani, S. N. Bhatia, Nat. Biotechnol. 2008, 26, 120. S. V. Murphy, J. D. Jackson, A. J. Atala, Sci. Rep. 2018, 8, 7413.
[200] P. G. Miller, M. L. Shuler, Biotechnol. Bioeng. 2016, 113, 2213. [226] Y. Y. Choi, B. G. Chung, D. H. Lee, A. Khademhosseini, J. H. Kim,
[201] C. D. Edington, W. L. K. Chen, E. Geishecker, T. Kassis, S. H. Lee, Biomaterials 2010, 31, 4296.
L. R. Soenksen, B. M. Bhushan, D. Freake, J. Kirschner, C. Maass, [227] Y. T. Dingle, M. E. Boutin, A. M. Chirila, L. L. Livi, N. R. Labriola,
N. Tsamandouras, J. Valdez, C. D. Cook, T. Parent, S. Snyder, J. Yu, L. M. Jakubek, J. R. Morgan, E. M. Darling, J. A. Kauer,
E. Suter, M. Shockley, J. Velazquez, J. J. Velazquez, L. Stockdale, D. Hoffman-Kim, Tissue Eng., Part C 2015, 21, 1274.
J. P. Papps, I. Lee, N. Vann, M. Gamboa, M. E. LaBarge, Z. Zhong, [228] Y. J. Choi, J. Park, S. H. Lee, Biomaterials 2013, 34, 2938.
X. Wang, L. A. Boyer, D. A. Lauffenburger, R. L. Carrier, et al., Sci. [229] J. Park, B. K. Lee, G. S. Jeong, J. K. Hyun, C. J. Lee, S. H. Lee, Lab
Rep. 2018, 8, 4530. Chip 2015, 15, 141.
[202] B. Zhang, A. Korolj, B. F. L. Lai, M. Radisic, Nat. Rev. Mater. 2018, [230] X. Yin, B. E. Mead, H. Safaee, R. Langer, J. M. Karp, O. Levy, Cell
3, 257. Stem Cell 2016, 18, 25.
[203] S. N. Bhatia, D. E. Ingber, Nat. Biotechnol. 2014, 32, 760. [231] Y. Yi, J. Park, J. Lim, C. J. Lee, S. H. Lee, Trends Biotechnol. 2015,
[204] D. C. Duffy, J. C. McDonald, O. J. A. Schueller, G. M. Whitesides, 33, 762.
Anal. Chem. 1998, 70, 4974. [232] H. L. Song, S. Shim, D. H. Kim, S. H. Won, S. Joo, S. Kim,
[205] M. A. Mofazzal Jahromi, A. Abdoli, M. Rahmanian, H. Bardania, N. L. Jeon, S. Y. Yoon, Ann. Neurol. 2014, 75, 88.
M. Bayandori, S. M. Moosavi Basri, A. Kalbasi, A. R. Aref, [233] W. W. Poon, M. Blurton-Jones, C. H. Tu, L. M. Feinberg,
M. Karimi, M. R. Hamblin, Mol. Neurobiol. 2019, 56, 8489. M. A. Chabrier, J. W. Harris, N. L. Jeon, C. W. Cotman, Neurobiol.
[206] R. Booth, H. Kim, Lab Chip 2012, 12, 1784. Aging 2011, 32, 821.
[207] B. Prabhakarpandian, M. C. Shen, J. B. Nichols, I. R. Mills, [234] B. Deleglise, S. Magnifico, E. Duplus, P. Vaur, V. Soubeyre,
M. Sidoryk-Wegrzynowicz, M. Aschner, K. Pant, Lab Chip 2013, 13, M. Belle, M. Vignes, J. L. Viovy, E. Jacotot, J. M. Peyrin, B. Brugg,
1093. Acta Neuropathol. Commun. 2014, 2, 145.
[208] L. M. Griep, F. Wolbers, B. De Wagenaar, P. M. Ter Braak, B. B. Weksler, [235] H. Cho, T. Hashimoto, E. Wong, Y. Hori, L. B. Wood, L. Zhao,
I. A. Romero, P. O. Couraud, I. Vermes, A. D. Van Der Meer, K. M. Haigis, B. T. Hyman, D. Irimia, Sci. Rep. 2013, 3, 1823.
A. Van Den Berg, Biomed. Microdevices 2013, 15, 145. [236] S. Dujardin, K. Lecolle, R. Caillierez, S. Begard, N. Zommer,
[209] G. Adriani, D. Ma, A. Pavesi, R. D. Kamm, E. L. Goh, Lab Chip C. Lachaud, S. Carrier, N. Dufour, G. Auregan, J. Winderickx,
2017, 17, 448. P. Hantraye, N. Deglon, M. Colin, L. Buee, Acta Neuropathol.
[210] T. E. Park, N. Mustafaoglu, A. Herland, R. Hasselkus, R. Mannix, Commun. 2014, 2, 14.
E. A. FitzGerald, R. Prantil-Baun, A. Watters, O. Henry, M. Benz, [237] J. W. Wu, S. A. Hussaini, I. M. Bastille, G. A. Rodriguez, A. Mrejeru,
H. Sanchez, H. J. McCrea, L. C. Goumnerova, H. W. Song, K. Rilett, D. W. Sanders, C. Cook, H. Fu, R. A. Boonen, M. Herman,
S. P. Palecek, E. Shusta, D. E. Ingber, Nat. Commun. 2019, 10, 2621. E. Nahmani, S. Emrani, Y. H. Figueroa, M. I. Diamond,
[211] J. A. Brown, V. Pensabene, D. A. Markov, V. Allwardt, M. D. Neely, C. L. Clelland, S. Wray, K. E. Duff, Nat. Neurosci. 2016, 19, 1085.
M. Shi, C. M. Britt, O. S. Hoilett, Q. Yang, B. M. Brewer, [238] J. Park, I. Wetzel, I. Marriott, D. Dreau, C. D’Avanzo, D. Y. Kim,
P. C. Samson, L. J. McCawley, J. M. May, D. J. Webb, D. Li, R. E. Tanzi, H. Cho, Nat. Neurosci. 2018, 21, 941.
A. B. Bowman, R. S. Reiserer, J. P. Wikswo, Biomicrofluidics 2015, [239] J. T. Fernandes, O. Chutna, V. Chu, J. P. Conde, T. F. Outeiro, Front.
9, 054124. Neurosci. 2016, 10, 511.
[212] G. D. Vatine, R. Barrile, M. J. Workman, S. Sances, B. K. Barriga, [240] S. Bolognin, M. Fossepre, X. Qing, J. Jarazo, J. Scancar,
M. Rahnama, S. Barthakur, M. Kasendra, C. Lucchesi, J. Kerns, E. L. Moreno, S. L. Nickels, K. Wasner, N. Ouzren, J. Walter,
N. Wen, W. R. Spivia, Z. Chen, J. Van Eyk, C. N. Svendsen, Cell A. Grunewald, E. Glaab, L. Salamanca, R. M. T. Fleming,
Stem Cell 2019, 24, 995. P. M. A. Antony, J. C. Schwamborn, Adv. Sci. 2019, 6, 1800927.
[213] Y. I. Wang, H. E. Abaci, M. L. Shuler, Biotechnol. Bioeng. 2017, 114, 184. [241] K. I. W. Kane, E. L. Moreno, S. Hachi, M. Walter, J. Jarazo,
[214] N. R. Wevers, D. G. Kasi, T. Gray, K. J. Wilschut, B. Smith, M. A. P. Oliveira, T. Hankemeier, P. Vulto, J. C. Schwamborn,
R. van Vught, F. Shimizu, Y. Sano, T. Kanda, G. Marsh, M. Thoma, R. M. T. Fleming, Sci. Rep. 2019, 9, 1796.
S. J. Trietsch, P. Vulto, H. L. Lanz, B. Obermeier, Fluids Barriers [242] M. Logun, W. Zhao, L. Mao, L. Karumbaiah, Adv. Biosyst. 2018, 2,
CNS 2018, 15, 23. 1700221.
[215] K. Ronaldson-Bouchard, G. Vunjak-Novakovic, Cell Stem Cell 2018, [243] J. Wang, D. Tham, W. Wei, Y. S. Shin, C. Ma, H. Ahmad, Q. Shi,
22, 310. J. Yu, R. D. Levine, J. R. Heath, Nano Lett. 2012, 12, 6101.

Adv. Funct. Mater. 2020, 1909146 1909146  (24 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[244] J. Sun, M. D. Masterman-Smith, N. A. Graham, J. Jiao, [270] G. Palazzolo, M. Moroni, A. Soloperto, G. Aletti, G. Naldi,
J. Mottahedeh, D. R. Laks, M. Ohashi, J. DeJesus, K. Kamei, M. Vassalli, T. Nieus, F. Difato, Sci. Rep. 2017, 7, 8499.
K. B. Lee, H. Wang, Z. T. Yu, Y. T. Lu, S. Hou, K. Li, M. Liu, [271] A. Marom, S. K. Mahto, E. Shor, J. Tenenbaum-Katan, J. Sznitman,
N. Zhang, S. Wang, B. Angenieux, E. Panosyan, E. R. Samuels, S. Shoham, Adv. Healthcare Mater. 2015, 4, 1478.
J. Park, D. Williams, V. Konkankit, D. Nathanson, R. M. van Dam, [272] J. Akerboom, N. Carreras Calderon, L. Tian, S. Wabnig, M. Prigge,
M. E. Phelps, H. Wu, L. M. Liau, P. S. Mischel, et al., Cancer Res. J. Tolo, A. Gordus, M. B. Orger, K. E. Severi, J. J. Macklin, R. Patel,
2010, 70, 6128. S. R. Pulver, T. J. Wardill, E. Fischer, C. Schuler, T. W. Chen,
[245] A. Pathak, S. Kumar, Proc. Natl. Acad. Sci. USA 2012, 109, 10334. K. S. Sarkisyan, J. S. Marvin, C. I. Bargmann, D. S. Kim, S. Kugler,
[246] J. Han, S. Oh, H. H. Hoang, D. T. T. Nguyen, W. Lim, T. H. Shin, L. Lagnado, P. Hegemann, A. Gottschalk, E. R. Schreiter, L. L. Looger,
G. Lee, S. Park, J. Ind. Eng. Chem. 2018, 62, 352. Front. Mol. Neurosci. 2013, 6, 2.
[247] J. M. Ayuso, R. Monge, A. Martinez-Gonzalez, M. Virumbrales-Munoz, [273] E. S. Boyden, F. Zhang, E. Bamberg, G. Nagel, K. Deisseroth, Nat.
G. A. Llamazares, J. Berganzo, A. Hernandez-Lain, J. Santolaria, Neurosci. 2005, 8, 1263.
M. Doblare, C. Hubert, J. N. Rich, P. Sanchez-Gomez, V. M. Perez-Garcia, [274] F. Zhang, L. P. Wang, M. Brauner, J. F. Liewald, K. Kay, N. Watzke,
I. Ochoa, L. J. Fernandez, Neuro-oncology 2017, 19, 503. P. G. Wood, E. Bamberg, G. Nagel, A. Gottschalk, K. Deisseroth,
[248] Y. Chonan, S. Taki, O. Sampetrean, H. Saya, R. Sudo, Integr. Biol. Nature 2007, 446, 633.
2017, 9, 762. [275] X. Han, E. S. Boyden, PLoS One 2007, 2, e299.
[249] D. Truong, R. Fiorelli, E. S. Barrientos, E. L. Melendez, N. Sanai, [276] R. Renault, N. Sukenik, S. Descroix, L. Malaquin, J. L. Viovy,
S. Mehta, M. Nikkhah, Biomaterials 2019, 198, 63. J. M. Peyrin, S. Bottani, P. Monceau, E. Moses, M. Vignes, PLoS
[250] Y. Fan, D. T. Nguyen, Y. Akay, F. Xu, M. Akay, Sci. Rep. 2016, 6, One 2015, 10, e0120680.
25062. [277] T. Kobayashi, M. Haruta, K. Sasagawa, M. Matsumata, K. Eizumi,
[251] M. Akay, J. Hite, N. G. Avci, Y. Fan, Y. Akay, G. Lu, J. J. Zhu, Sci. C. Kitsumoto, M. Motoyama, Y. Maezawa, Y. Ohta, T. Noda,
Rep. 2018, 8, 15423. T. Tokuda, Y. Ishikawa, J. Ohta, Sci. Rep. 2016, 6, 21247.
[252] H. Shao, J. Chung, K. Lee, L. Balaj, C. Min, B. S. Carter, F. H. Hochberg, [278] J. Delbeke, L. Hoffman, K. Mols, D. Braeken, D. Prodanov, Front.
X. O. Breakefield, H. Lee, R. Weissleder, Nat. Commun. 2015, 6, 6999. Neurosci. 2017, 11, 663.
[253] T. C. Chang, A. M. Mikheev, W. Huynh, R. J. Monnat, [279] D. R. Hochbaum, Y. Zhao, S. L. Farhi, N. Klapoetke, C. A. Werley,
R. C. Rostomily, A. Folch, Lab Chip 2014, 14, 4540. V. Kapoor, P. Zou, J. M. Kralj, D. Maclaurin, N. Smedemark-Margu
[254] J. Wu, G. Zheng, L. M. Lee, Lab Chip 2012, 12, 3566. lies, J. L. Saulnier, G. L. Boulting, C. Straub, Y. K. Cho, M. Melkonian,
[255] I. Albert-Smet, A. Marcos-Vidal, J. J. Vaquero, M. Desco, G. K. Wong, D. J. Harrison, V. N. Murthy, B. L. Sabatini, E. S. Boyden,
A. Munoz-Barrutia, J. Ripoll, Front. Neuroanat. 2019, 13, 1. R. E. Campbell, A. E. Cohen, Nat. Methods 2014, 11, 825.
[256] P. Paie, R. Martinez Vazquez, R. Osellame, F. Bragheri, A. Bassi, [280] G. Huang, F. Li, X. Zhao, Y. Ma, Y. Li, M. Lin, G. Jin, T. J. Lu,
Cytometry, Part A 2018, 93, 987. G. M. Genin, F. Xu, Chem. Rev. 2017, 117, 12764.
[257] Z. Geng, X. Zhang, Z. Fan, X. Lv, Y. Su, H. Chen, Sensors 2017, 17, [281] L. Yildirimer, Q. Zhang, S. Kuang, C. J. Cheung, K. A. Chu, Y. He,
2449. M. Yang, X. Zhao, Biofabrication 2019, 11, 032003.
[258] S. Cho, A. Islas-Robles, A. M. Nicolini, T. J. Monks, J. Y. Yoon, [282] A. R. Murphy, A. Laslett, C. M. O’Brien, N. R. Cameron, Acta
Biosens. Bioelectron. 2016, 86, 697. Biomater. 2017, 54, 1.
[259] F. Hofmann, H. Bading, J. Physiol. 2006, 99, 125. [283] E. R. Aurand, K. J. Lampe, K. B. Bjugstad, Neurosci. Res. 2012, 72, 199.
[260] E. Moutaux, B. Charlot, A. Genoux, F. Saudou, M. Cazorla, Lab [284] E. Ruoslahti, Glycobiology 1996, 6, 489.
Chip 2018, 18, 3425. [285] P. C. Georges, W. J. Miller, D. F. Meaney, E. S. Sawyer, P. A. Janmey,
[261] T. T. Kanagasabapathi, M. Franco, R. A. Barone, S. Martinoia, Biophys. J. 2006, 90, 3012.
W. J. Wadman, M. M. Decre, J. Neurosci. Methods 2013, 214, 1. [286] W. Ma, W. Fitzgerald, Q. Y. Liu, T. J. O’Shaughnessy, D. Maric,
[262] H. T. Hogberg, T. Sobanski, A. Novellino, M. Whelan, D. G. Weiss, H. J. Lin, D. L. Alkon, J. L. Barker, Exp. Neurol. 2004, 190, 276.
A. K. Bal-Price, Neurotoxicology 2011, 32, 158. [287] K. Watanabe, M. Nakamura, H. Okano, Y. Toyama, Restor. Neurol.
[263] B. Scelfo, M. Politi, F. Reniero, T. Palosaari, M. Whelan, Neurosci. 2007, 25, 109.
J. M. Zaldivar, Toxicology 2012, 299, 172. [288] K. Brannvall, K. Bergman, U. Wallenquist, S. Svahn, T. Bowden,
[264] L. Yla-Outinen, J. Heikkila, H. Skottman, R. Suuronen, J. Hilborn, K. Forsberg-Nilsson, J. Neurosci. Res. 2007, 85, 2138.
R. Aanismaa, S. Narkilahti, Front. Neuroeng. 2010, 3, 111. [289] Y. B. Lee, S. Polio, W. Lee, G. Dai, L. Menon, R. S. Carroll, S. S. Yoo,
[265] E. Defranchi, A. Novellino, M. Whelan, S. Vogel, T. Ramirez, B. van Exp. Neurol. 2010, 223, 645.
 Ravenzwaay, R. Landsiedel, Front. Neuroeng. 2011, 4, 6. [290] C. Deister, C. E. Schmidt, J. Neural Eng. 2006, 3, 172.
[266] A. Scott, K. Weir, C. Easton, W. Huynh, W. J. Moody, A. Folch, Lab [291] H. G. Sundararaghavan, G. A. Monteiro, B. L. Firestein,
Chip 2013, 13, 527. D. I. Shreiber, Biotechnol. Bioeng. 2009, 102, 632.
[267] Y. S. Zhang, J. Aleman, S. R. Shin, T. Kilic, D. Kim, S. A. Mousavi [292] C. Bouyer, P. Chen, S. Guven, T. T. Demirtas, T. J. Nieland,
Shaegh, S. Massa, R. Riahi, S. Chae, N. Hu, H. Avci, W. Zhang, F. Padilla, U. Demirci, Adv. Mater. 2016, 28, 161.
A. Silvestri, A. Sanati Nezhad, A. Manbohi, F. De Ferrari, A. Polini, [293] V. A. Kornev, E. A. Grebenik, A. B. Solovieva, R. I. Dmitriev,
G. Calzone, N. Shaikh, P. Alerasool, E. Budina, J. Kang, N. Bhise, P. S. Timashev, Comput. Struct. Biotechnol. J. 2018, 16, 488.
J. Ribas, A. Pourmand, A. Skardal, T. Shupe, C. E. Bishop, [294] X. Wang, J. He, Y. Wang, F. Z. Cui, Interface Focus 2012, 2, 278.
M. R. Dokmeci, A. Atala, A. Khademhosseini, Proc. Natl. Acad. Sci. [295] S. Hou, W. Tian, Q. Xu, F. Cui, J. Zhang, Q. Lu, C. Zhao, Neurosci-
USA 2017, 114, E2293. ence 2006, 137, 519.
[268] J. Nakai, M. Ohkura, K. Imoto, Nat. Biotechnol. 2001, 19, 137. [296] W. M. Tian, S. P. Hou, J. Ma, C. L. Zhang, Q. Y. Xu, I. S. Lee,
[269] J. Akerboom, T. W. Chen, T. J. Wardill, L. Tian, J. S. Marvin, H. D. Li, M. Spector, F. Z. Cui, Tissue Eng. 2005, 11, 513.
S. Mutlu, N. C. Calderon, F. Esposti, B. G. Borghuis, X. R. Sun, [297] S. Suri, C. E. Schmidt, Tissue Eng., Part A 2010, 16, 1703.
A. Gordus, M. B. Orger, R. Portugues, F. Engert, J. J. Macklin, [298] S. K. Seidlits, Z. Z. Khaing, R. R. Petersen, J. D. Nickels,
A. Filosa, A. Aggarwal, R. A. Kerr, R. Takagi, S. Kracun, J. E. Vanscoy, J. B. Shear, C. E. Schmidt, Biomaterials 2010, 31,
E. Shigetomi, B. S. Khakh, H. Baier, L. Lagnado, S. S. Wang, 3930.
C. I. Bargmann, B. E. Kimmel, V. Jayaraman, K. Svoboda, [299] M. Matyash, F. Despang, R. Mandal, D. Fiore, M. Gelinsky,
D. S. Kim, et al., J. Neurosci. 2012, 32, 13819. C. Ikonomidou, Tissue Eng., Part A 2012, 18, 55.

Adv. Funct. Mater. 2020, 1909146 1909146  (25 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[300] J. P. Frampton, M. R. Hynd, M. L. Shuler, W. Shain, Biomed. Mater. [331] Y. Lai, A. Asthana, K. Cheng, W. S. Kisaalita, PLoS One 2011, 6,
2011, 6, 015002. e26821.
[301] S. R. Moxon, N. J. Corbett, K. Fisher, G. Potjewyd, M. Domingos, [332] Y. Lai, K. Cheng, W. Kisaalita, PLoS One 2012, 7, e45074.
N. M. Hooper, Mater. Sci. Eng., C 2019, 104, 109904. [333] N. Li, Q. Zhang, S. Gao, Q. Song, R. Huang, L. Wang, L. Liu,
[302] A. Kunze, M. Giugliano, A. Valero, P. Renaud, Biomaterials 2011, J. Dai, M. Tang, G. Cheng, Sci. Rep. 2013, 3, 1604.
32, 2088. [334] A. Jakobsson, M. Ottosson, M. C. Zalis, D. O’Carroll,
[303] C. R. Kothapalli, R. D. Kamm, Biomaterials 2013, 34, 5995. U. E. Johansson, F. Johansson, Nanomed.: Nanotechnol., Biol.
[304] A. M. Hopkins, L. De Laporte, F. Tortelli, E. Spedden, C. Staii, Med. 2017, 13, 1563.
T. J. Atherton, J. A. Hubbell, D. L. Kaplan, Adv. Funct. Mater. 2013, [335] B. A. Harley, J. H. Leung, E. C. Silva, L. J. Gibson, Acta Biomater.
23, 5140. 2007, 3, 463.
[305] X. Li, E. Katsanevakis, X. Liu, N. Zhang, X. Wen, Prog. Polym. Sci. [336] M. G. Haugh, C. M. Murphy, F. J. O’Brien, Tissue Eng., Part C 2010,
2012, 37, 1105. 16, 887.
[306] M. J. Mahoney, K. S. Anseth, Biomaterials 2006, 27, 2265. [337] T. W. Wang, M. Spector, Acta Biomater. 2009, 5, 2371.
[307] J. W. Gunn, S. D. Turner, B. K. Mann, J. Biomed. Mater. Res., Part A [338] P. Z. Elias, M. Spector, J. Neurosci. Methods 2012, 209, 199.
2005, 72, 91. [339] H. Hosseinkhani, Y. Hiraoka, C. H. Li, Y. R. Chen, D. S. Yu,
[308] Z. N. Zhang, B. C. Freitas, H. Qian, J. Lux, A. Acab, C. A. Trujillo, P. D. Hong, K. L. Ou, ACS Chem. Neurosci. 2013, 4, 1229.
R. H. Herai, V. A. Nguyen Huu, J. H. Wen, S. Joshi-Barr, [340] M. Jurga, M. B. Dainiak, A. Sarnowska, A. Jablonska, A. Tripathi,
J. V. Karpiak, A. J. Engler, X. D. Fu, A. R. Muotri, A. Almutairi, Proc. F. M. Plieva, I. N. Savina, L. Strojek, H. Jungvid, A. Kumar,
Natl. Acad. Sci. USA 2016, 113, 3185. B. Lukomska, K. Domanska-Janik, N. Forraz, C. P. McGuckin,
[309] M. V. Tsurkan, K. Chwalek, S. Prokoph, A. Zieris, K. R. Levental, Biomaterials 2011, 32, 3423.
U. Freudenberg, C. Werner, Adv. Mater. 2013, 25, 2606. [341] M. D. Tang-Schomer, J. D. White, L. W. Tien, L. I. Schmitt,
[310] L. M. Yu, K. Kazazian, M. S. Shoichet, J. Biomed. Mater. Res., Part T. M. Valentin, D. J. Graziano, A. M. Hopkins, F. G. Omenetto,
A 2007, 82, 243. P. G. Haydon, D. L. Kaplan, Proc. Natl. Acad. Sci. USA 2014, 111,
[311] H. Li, A. Wijekoon, N. D. Leipzig, PLoS One 2012, 7, e48824. 13811.
[312] H. Li, A. Wijekoon, N. D. Leipzig, Ann. Biomed. Eng. 2014, 42, [342] V. Liaudanskaya, D. Jgamadze, A. N. Berk, D. J. Bischoff, B. J. Gu,
1456. H. Hawks-Mayer, M. J. Whalen, H. I. Chen, D. L. Kaplan, Biomate-
[313] T. C. Holmes, S. de Lacalle, X. Su, G. Liu, A. Rich, S. Zhang, Proc. rials 2019, 192, 510.
Natl. Acad. Sci. USA 2000, 97, 6728. [343] D. Zhang, M. Pekkanen-Mattila, M. Shahsavani, A. Falk,
[314] Z. X. Zhang, Q. X. Zheng, Y. C. Wu, D. J. Hao, Biotechnol. A. I. Teixeira, A. Herland, Biomaterials 2014, 35, 1420.
Bioprocess Eng. 2010, 15, 545. [344] M. N. Labour, S. Vigier, D. Lerner, A. Marcilhac, E. Belamie, Acta
[315] S. Ortinau, J. Schmich, S. Block, A. Liedmann, L. Jonas, Biomater. 2016, 37, 38.
D. G. Weiss, C. A. Helm, A. Rolfs, M. J. Frech, Biomed. Eng. Online [345] C. Papadimitriou, H. Celikkaya, M. I. Cosacak, V. Mashkaryan,
2010, 9, 70. L. Bray, P. Bhattarai, K. Brandt, H. Hollak, X. Chen, S. He,
[316] T. Y. Cheng, M. H. Chen, W. H. Chang, M. Y. Huang, T. W. Wang, C. L. Antos, W. Lin, A. K. Thomas, A. Dahl, T. Kurth, J. Friedrichs,
Biomaterials 2013, 34, 2005. Y. Zhang, U. Freudenberg, C. Werner, C. Kizil, Dev. Cell 2018, 46,
[317] Y. Sun, W. Li, X. Wu, N. Zhang, Y. Zhang, S. Ouyang, X. Song, 85.
X. Fang, R. Seeram, W. Xue, L. He, W. Wu, ACS Appl. Mater. Inter- [346] W. L. Cantley, C. Du, S. Lomoio, T. Depalma, E. Peirent,
faces 2016, 8, 2348. D. Kleinknecht, M. Hunter, M. D. Tang-Schomer, G. Tesco,
[318] A. Caprini, D. Silva, I. Zanoni, C. Cunha, C. Volonte, A. Vescovi, D. L. Kaplan, ACS Biomater. Sci. Eng. 2018, 4, 4278.
F. Gelain, New Biotechnol. 2013, 30, 552. [347] P. M. D. Watson, E. Kavanagh, G. Allenby, M. Vassey, SLAS
[319] R. Pugliese, F. Fontana, A. Marchini, F. Gelain, Acta Biomater. Discovery 2017, 22, 583.
2018, 66, 258. [348] W. Xiao, A. Sohrabi, S. K. Seidlits, Future Sci. OA 2017, 3, Fso189.
[320] A. Marchini, A. Raspa, R. Pugliese, M. A. El Malek, V. Pastori, [349] B. Ananthanarayanan, Y. Kim, S. Kumar, Biomaterials 2011, 32,
M. Lecchi, A. L. Vescovi, F. Gelain, Proc. Natl. Acad. Sci. USA 2019, 7913.
116, 7483. [350] C. Wang, X. Tong, F. Yang, Mol. Pharmaceutics 2014, 11, 2115.
[321] M. Hospodiuk, M. Dey, D. Sosnoski, I. T. Ozbolat, Biotechnol. Adv. [351] S. J. Florczyk, K. Wang, S. Jana, D. L. Wood, S. K. Sytsma, J. Sham,
2017, 35, 217. F. M. Kievit, M. Zhang, Biomaterials 2013, 34, 10143.
[322] S. Knowlton, S. Anand, T. Shah, S. Tasoglu, Trends Neurosci. 2018, [352] D. Sood, M. Tang-Schomer, D. Pouli, C. Mizzoni, N. Raia, A. Tai,
41, 31. K. Arkun, J. Wu, L. D. Black 3rd, B. Scheffler, I. Georgakoudi,
[323] T. Xu, C. A. Gregory, P. Molnar, X. Cui, S. Jalota, S. B. Bhaduri, D. A. Steindler, D. L. Kaplan, Nat. Commun. 2019, 10, 4529.
T. Boland, Biomaterials 2006, 27, 3580. [353] S. Knowlton, S. Onal, C. H. Yu, J. J. Zhao, S. Tasoglu, Trends
[324] W. Lee, J. Pinckney, V. Lee, J. H. Lee, K. Fischer, S. Polio, J. K. Park, Biotechnol. 2015, 33, 504.
S. S. Yoo, NeuroReport 2009, 20, 798. [354] X. Dai, C. Ma, Q. Lan, T. Xu, Biofabrication 2016, 8, 045005.
[325] F. Y. Hsieh, H. H. Lin, S. H. Hsu, Biomaterials 2015, 71, 48. [355] S. Pedron, E. Becka, B. A. Harley, Adv. Mater. 2015, 27, 1567.
[326] X. Zhou, H. Cui, M. Nowicki, S. Miao, S. J. Lee, F. Masood, [356] M. A. Heinrich, R. Bansal, T. Lammers, Y. S. Zhang, R. Michel
B. T. Harris, L. G. Zhang, ACS Appl. Mater. Interfaces 2018, 10, 8993. Schiffelers, J. Prakash, Adv. Mater. 2019, 31, 1806590.
[327] R. Lozano, L. Stevens, B. C. Thompson, K. J. Gilmore, [357] S. Caragher, A. J. Chalmers, N. Gomez-Roman, Cancers 2019, 11, 44.
R. Gorkin 3rd, E. M. Stewart, M. in het Panhuis, [358] N. Gomez-Roman, K. Stevenson, L. Gilmour, G. Hamilton,
M. Romero-Ortega, G. G. Wallace, Biomaterials 2015, 67, 264. A. J. Chalmers, Neuro-oncology 2017, 19, 229.
[328] Q. Gu, E. Tomaskovic-Crook, R. Lozano, Y. Chen, R. M. Kapsa, [359] E. Bar-Kochba, M. T. Scimone, J. B. Estrada, C. Franck, Sci. Rep.
Q. Zhou, G. G. Wallace, J. M. Crook, Adv. Healthcare Mater. 2016, 2016, 6, 30550.
5, 1429. [360] M. P. Schwartz, Z. Hou, N. E. Propson, J. Zhang, C. J. Engstrom,
[329] Q. Gu, E. Tomaskovic-Crook, G. G. Wallace, J. M. Crook, Adv. V. Santos Costa, P. Jiang, B. K. Nguyen, J. M. Bolin, W. Daly,
Healthcare Mater. 2017, 6, 1700175. Y. Wang, R. Stewart, C. D. Page, W. L. Murphy, J. A. Thomson,
[330] K. Cheng, Y. Lai, W. S. Kisaalita, Biomaterials 2008, 29, 2802. Proc. Natl. Acad. Sci. USA 2015, 112, 12516.

Adv. Funct. Mater. 2020, 1909146 1909146  (26 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.afm-journal.de

[361] A. Ranga, M. Girgin, A. Meinhardt, D. Eberle, M. Caiazzo, [379] J. D. Miller, Y. M. Ganat, S. Kishinevsky, R. L. Bowman, B. Liu,
E. M. Tanaka, M. P. Lutolf, Proc. Natl. Acad. Sci. USA 2016, 113, E. Y. Tu, P. K. Mandal, E. Vera, J. W. Shim, S. Kriks, T. Taldone,
E6831. N. Fusaki, M. J. Tomishima, D. Krainc, T. A. Milner, D. J. Rossi,
[362] M. J. Kratochvil, A. J. Seymour, T. L. Li, S. P. Paşca, C. J. Kuo, L. Studer, Cell Stem Cell 2013, 13, 691.
S. C. Heilshorn, Nat. Rev. Mater. 2019, 4, 606. [380] J. Mertens, A. C. M. Paquola, M. Ku, E. Hatch, L. Bohnke,
[363] G. Y. Cederquist, J. J. Asciolla, J. Tchieu, R. M. Walsh, D. Cornacchia, S. Ladjevardi, S. McGrath, B. Campbell, H. Lee, J. R. Herdy,
M. D. Resh, L. Studer, Nat. Biotechnol. 2019, 37, 436. J. T. Goncalves, T. Toda, Y. Kim, J. Winkler, J. Yao, M. W. Hetzer,
[364] G. T. Knight, B. F. Lundin, N. Iyer, L. M. Ashton, W. A. Sethares, F. H. Gage, Cell Stem Cell 2015, 17, 705.
R. M. Willett, R. S. Ashton, eLife 2018, 7, e37549. [381] R. R. White, J. Vijg, Mol. Cell 2016, 63, 729.
[365] J. Lam, N. F. Truong, T. Segura, Acta Biomater. 2014, 10, 1571. [382] D. D. Raj, D. Jaarsma, I. R. Holtman, M. Olah, F. M. Ferreira,
[366] R. Boni, A. Ali, A. Shavandi, A. N. Clarkson, J. Biomed. Sci. 2018, W. Schaafsma, N. Brouwer, M. M. Meijer, M. C. de Waard,
25, 90. I. van der Pluijm, R. Brandt, K. L. Kreft, J. D. Laman, G. de Haan,
[367] S. Ilkhanizadeh, A. I. Teixeira, O. Hermanson, Biomaterials 2007, K. P. Biber, J. H. Hoeijmakers, B. J. Eggen, H. W. Boddeke, Neuro-
28, 3936. biol. Aging 2014, 35, 2147.
[368] G. Zhao, H. Qing, G. Huang, G. M. Genin, T. J. Lu, Z. Luo, F. Xu, [383] M. C. de Waard, I. van der Pluijm, N. Zuiderveen Borgesius,
X. Zhang, NPG Asia Mater. 2018, 10, 982. L. H. Comley, E. D. Haasdijk, Y. Rijksen, Y. Ridwan, G. Zondag,
[369] Y. H. Zhao, C. M. Niu, J. Q. Shi, Y. Y. Wang, Y. M. Yang, H. B. Wang, J. H. Hoeijmakers, Y. Elgersma, T. H. Gillingwater, D. Jaarsma,
Neural Regener. Res. 2018, 13, 1455. Acta Neuropathol. 2010, 120, 461.
[370] S. L. Giandomenico, S. B. Mierau, G. M. Gibbons, L. M. D. Wenger, [384] N. Z. Borgesius, M. C. de Waard, I. van der Pluijm, A. Omrani,
L. Masullo, T. Sit, M. Sutcliffe, J. Boulanger, M. Tripodi, E. Derivery, G. C. Zondag, G. T. van der Horst, D. W. Melton, J. H. Hoeijmakers,
O. Paulsen, A. Lakatos, M. A. Lancaster, Nat. Neurosci. 2019, 22, D. Jaarsma, Y. Elgersma, J. Neurosci. 2011, 31, 12543.
669. [385] M. J. Vegh, M. C. de Waard, I. van der Pluijm, Y. Ridwan,
[371] N. Ashammakhi, S. Ahadian, C. Xu, H. Montazerian, H. Ko, M. J. Sassen, P. van Nierop, R. C. van der Schors, K. W. Li,
R. Nasiri, N. Barros, A. Khademhosseini, Mater. Today Bio 2019, 1, J. H. Hoeijmakers, A. B. Smit, R. E. van Kesteren, J. Proteome Res.
100008. 2012, 11, 1855.
[372] W. Ma, T. Tavakoli, S. Chen, D. Maric, J. L. Liu, T. J. O’Shaughnessy, [386] E. L. de Graaf, W. P. Vermeij, M. C. de Waard, Y. Rijksen, I. van
J. L. Barker, Tissue Eng., Part A 2008, 14, 1673. der Pluijm, C. C. Hoogenraad, J. H. Hoeijmakers, A. F. Altelaar,
[373] M. T. Raimondi, D. Albani, C. Giordano, Trends Mol. Med. 2019, 25, 737. A. J. Heck, Mol. Cell. Proteomics 2013, 12, 1350.
[374] M. Kato-Negishi, Y. Morimoto, H. Onoe, S. Takeuchi, Adv. Health- [387] E. Vera, N. Bosco, L. Studer, Cell Rep. 2016, 17, 1184.
care Mater. 2013, 2, 1564. [388] Y. Lai, W. S. Kisaalita, J. Lab. Autom. 2012, 17, 284.
[375] M. F. Hasan, Y. Berdichevsky, Micromachines 2016, 7, 157. [389] K. Domansky, D. C. Leslie, J. McKinney, J. P. Fraser, J. D. Sliz,
[376] K. Ndyabawe, W. S. Kisaalita, Drug Discovery Today 2019, 24, 1725. T. Hamkins-Indik, G. A. Hamilton, A. Bahinski, D. E. Ingber, Lab
[377] A. De Sandre-Giovannoli, R. Bernard, P. Cau, C. Navarro, J. Amiel, Chip 2013, 13, 3956.
I. Boccaccio, S. Lyonnet, C. L. Stewart, A. Munnich, M. Le Merrer, [390] C. S. Bjornsson, G. Lin, Y. Al-Kofahi, A. Narayanaswamy, K. L. Smith,
N. Lévy, Science 2003, 300, 2055. W. Shain, B. Roysam, J. Neurosci. Methods 2008, 170, 165.
[378] M. Eriksson, W. T. Brown, L. B. Gordon, M. W. Glynn, J. Singer, [391] H. G. Jahnke, A. Braesigk, T. G. Mack, S. Ponick, F. Striggow,
L. Scott, M. R. Erdos, C. M. Robbins, T. Y. Moses, P. Berglund, A. A. Robitzki, Biosens. Bioelectron. 2012, 32, 250.
A. Dutra, E. Pak, S. Durkin, A. B. Csoka, M. Boehnke, T. W. Glover, [392] H. Dana, A. Marom, S. Paluch, R. Dvorkin, I. Brosh, S. Shoham,
F. S. Collins, Nature 2003, 423, 293. Nat. Commun. 2014, 5, 3997.

Adv. Funct. Mater. 2020, 1909146 1909146  (27 of 27) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like