ArticlePDF AvailableLiterature Review

Blood-Brain Barrier Breakdown in Neuroinflammation: Current In Vitro Models

Authors:

Abstract

The blood-brain barrier, which is formed by tightly interconnected microvascular endothe-lial cells, separates the brain from the peripheral circulation. Together with other central nervous system-resident cell types, including pericytes and astrocytes, the blood-brain barrier forms the neurovascular unit. Upon neuroinflammation, this barrier becomes leaky, allowing molecules and cells to enter the brain and to potentially harm the tissue of the central nervous system. Despite the significance of animal models in research, they may not always adequately reflect human patho-physiology. Therefore, human models are needed. This review will provide an overview of the blood-brain barrier in terms of both health and disease. It will describe all key elements of the in vitro models and will explore how different compositions can be utilized to effectively model a variety of neuroinflammatory conditions. Furthermore, it will explore the existing types of models that are used in basic research to study the respective pathologies thus far.
Citation: Brandl, S.; Reindl, M.
Blood–Brain Barrier Breakdown in
Neuroinflammation: Current In Vitro
Models. Int. J. Mol. Sci. 2023,24,
12699. https://doi.org/10.3390/
ijms241612699
Academic Editor: Monique F. Stins
Received: 21 June 2023
Revised: 7 August 2023
Accepted: 8 August 2023
Published: 11 August 2023
Copyright: © 2023 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
International Journal of
Molecular Sciences
Review
Blood–Brain Barrier Breakdown in Neuroinflammation:
Current In Vitro Models
Sarah Brandl and Markus Reindl *
Clinical Department of Neurology, Medical University of Innsbruck, 6020 Innsbruck, Austria;
sarah.brandl@i-med.ac.at
*Correspondence: markus.reindl@i-med.ac.at
Abstract:
The blood–brain barrier, which is formed by tightly interconnected microvascular endothe-
lial cells, separates the brain from the peripheral circulation. Together with other central nervous
system-resident cell types, including pericytes and astrocytes, the blood–brain barrier forms the
neurovascular unit. Upon neuroinflammation, this barrier becomes leaky, allowing molecules and
cells to enter the brain and to potentially harm the tissue of the central nervous system. Despite the
significance of animal models in research, they may not always adequately reflect human patho-
physiology. Therefore, human models are needed. This review will provide an overview of the
blood–brain barrier in terms of both health and disease. It will describe all key elements of the in vitro
models and will explore how different compositions can be utilized to effectively model a variety
of neuroinflammatory conditions. Furthermore, it will explore the existing types of models that are
used in basic research to study the respective pathologies thus far.
Keywords: blood–brain barrier; neurovascular unit; neuroinflammation; in vitro models
1. Introduction
Typically, 98% of small-molecule drugs, and close to 100% of large-molecule drugs, fail
to enter the central nervous system (CNS) through the tight barrier of endothelial cells [
1
].
Moreover, approximately 30% of all the drugs that are specifically developed for treating
CNS diseases encounter failure in penetrating the endothelial cell layer when attempting
to access the CNS [
2
]. However, under certain conditions, this physical barrier becomes
permeable. This allows both beneficial and detrimental cells, as well as substances to enter
the CNS tissue. Neuronal and glial injury arise due to the inflammatory environment
that results from the breakdown of the usually highly effective blood–brain barrier (BBB).
Studies investigating the BBB in different neurological disorders in vivo have only been partly
successful so far. These animal models did not always reflect the human (patho-) physiology,
and the results of animal studies cannot always be replicated in humans. To circumvent this
problem, but also to reduce and replace animal experiments, human tissue in vitro models are
still urgently needed. This review provides an overview of the current approaches in terms of
modeling the neuroinflammation of the BBB in different neurological and psychiatric diseases.
2. The Neurovascular Unit in Health
There are different definitions of the vascular BBB: they vary from just the brain
microvascular endothelial cells (BMECs) with pericytes in the basement membrane (BM),
astrocyte endfeet enwrapping the capillaries, and other cells signaling the formation of
barrier features, but not themselves participating in the physical barrier, to the inclusion of
the BM and glycocalyx. In contrast, the neurovascular unit (NVU) has become a collective
term for all cell types that are involved in BBB integrity, including microglia and neurons.
Int. J. Mol. Sci. 2023,24, 12699. https://doi.org/10.3390/ijms241612699 https://www.mdpi.com/journal/ijms
Int. J. Mol. Sci. 2023,24, 12699 2 of 40
2.1. Microvascular Endothelial Cells
BMECs are the key regulators of the brain microenvironment, and the most important
component of the BBB. They cover all brain microcapillaries with a total surface of between
12 and 18 m
2
in an adult human [
3
]. Moreover, BMECs express crucial factors regulating
the permeability of the BBB. Microvascular heterogeneity results from the diverse functions
that the cell layer needs to exhibit in different tissues. For example, the glomerular endothe-
lium of the kidney is fenestrated with intracellular pores for rapid filtration, absorption,
and secretion, whereas the sinusoidal endothelium of the spleen or liver is discontinuous
with larger pores for large solute exchanges at higher rates [
4
,
5
]. In contrast, BMECs exhibit
a continuous, tight layer with the help of gap-, adherens-, and tight junctions (TJs) that
bring them into proximity and limit paracellular transport across this layer [
6
,
7
]. Adherens
junctions consist of transmembrane cadherin–cadherin complexes between adjacent cells,
and they are responsible for the adhesion between cells, catenins, and scaffolding proteins
in the cytoplasm. Vascular endothelial-cadherins mediate the connection between two
BMECs, whereas N-cadherins establish the connections with pericytes [
8
]. Gap junctions
are composed of tissue-specific connexins that form hexamers at the membrane. The align-
ment of two neighboring endothelial cell–surface hemichannels allows for intercellular
communication through the exchange of ions and small molecules between BMECs [
7
,
9
].
TJs form the physical separation between the bloodstream and the CNS tissue. They
consist of the transmembrane proteins occludin, claudins (especially claudin-5), and junc-
tional adhesion molecules [
10
,
11
]. Their extracellular domains are in close interaction with
their counterparts on neighboring cells, thereby narrowing the paracellular cleft [
12
]. The
carboxyl-terminal of these transmembrane proteins interact with the actin cytoskeleton and
scaffolding proteins, such as zonula occludens-1 and -2 (ZO-1, -2). ZO-1 also establishes
interactions between adherens- and gap junction proteins [7].
Many reports from the literature state that only small lipophilic molecules with a
molecular weight below 400–600 Da and with a few (<8) hydrogen bonds, as well as gases,
can diffuse freely across the BBB. However, there is evidence that there is no distinct cutoff,
but rather a molecular weight penalty. There are certain molecules with a higher molec-
ular weight that can diffuse through the BBB by lipid solubility, the largest known being
cytokine-induced neutrophil chemoattractant-1 that has about 7 kDa. Furthermore, some
substrate classes, such as anti-helminthics, (opiate) peptides, and their analogs, can also
cross the BBB by diffusion despite higher molecular weights [
2
,
13
,
14
]. Additionally, diverse
transcellular transport systems for other molecules ensure a physiological brain pH and
metabolism [
15
]. Nutrients, such as glucose and amino acids, need solute carrier-mediated
transporters to enter the CNS. Many of them have a specific direction and substrate pre-
cision. Hormones, organic anions and cations, amines, nucleotides, vitamins, and fatty
acids are also carrier-mediated transported [
16
]. Receptors mediate the transport of pro-
tein ligands, such as amyloid-beta (A
β
), transferrin, insulin, and apolipoprotein E [
17
,
18
].
Ions are transported across the endothelium via adenosine triphosphatases, uniporters,
exchangers, and symporters [
15
]. The active efflux of drugs (or drug conjugates), xenobi-
otics, or nucleosides is mediated by adenosine triphosphate-binding cassette transporters.
The most prominent ones to name are P-glycoprotein (P-gp) and the multi-drug resistance
protein-1 [
6
]. To conclude, many proteins and other macromolecules cannot cross the BBB,
and they are held back in the bloodstream under physiological conditions.
Basolateral BMECs are embedded in the BM. This extracellular matrix (ECM) consists
primarily of structural elements, especially collagen type IV; specialized proteins such as
laminins, nidogen, fibronectin, and proteoglycans like perlecan; and agrin [
19
]. Different
α
-
and
β
-integrin receptors, which form transmembrane heterodimers, provide the functional
link between the BMEC cytoskeleton and the ECM. Additionally, the BM has an impact on
the integrity of endothelial junctions [19,20].
Toward the vascular lumen, BMECs express the glycocalyx, which is a thin layer of
a villiform substance [
21
]. Its major components are proteoglycan protein polymers and
glycosaminoglycan chains, including heparan sulfate, chondroitin sulfate, hyaluronic acid,
Int. J. Mol. Sci. 2023,24, 12699 3 of 40
and their associated binding proteins [
22
]. The glycocalyx is important for many of the
physiological functions of the BBB. Among them, it maintains the low permeability of the
BBB, prevents inflammation triggers, and the coagulation response [
21
]. Furthermore, it
was shown to sense changes in the shear force of the blood flow, subsequently inducing the
release of endogenous vasoactive mediators [
23
]. As it is negatively charged, the glycocalyx
forms an electrostatic barrier for negatively charged molecules, proteins, and plasma
cells [
24
]. During inflammation, the glycocalyx sheds off the BMECs to enable leukocyte
binding to vascular cell adhesion molecules [
21
]. Furthermore, glycocalyx degradation
decreases the physical barrier, enhances the permeability, promotes inflammation by a direct
interaction of the BMECs with plasma components or blood cells, as well as interferes with
various receptor functions, such as lipid mediators [
24
,
25
]. Alterations in the glycocalyx
have been shown to affect the BBB integrity, which could promote the development of a
broad range of neurological diseases [26].
2.2. Pericytes
Pericytes are present in most of the non-epithelial tissues around vessels. However,
they are most abundant in the CNS, especially in the retina, where they cover approximately
30% of the vessel surface with varying frequencies depending on their location [
27
,
28
].
They are embedded in the BM, which is where they are in close association with BMECs at
a distance of less than 20 nm. The membrane facing the microvessels expresses N-cadherin
and connexins, which bring the two cell types into a closer proximity. Through their
close interaction, the two cell types can exchange ions, second messengers, metabolites,
and ribonucleic acids [
8
]. The distinct functions of pericytes have been discussed rigor-
ously, with contractile abilities that resemble those of smooth muscle cells being the most
significant function [
29
]. Furthermore, they play an important role for the BBB regarding
integrity, thereby supporting angiogenesis and microvascular stability [
30
]. This becomes
especially evident in the platelet-derived growth factor receptor-
β
deficient mice, which
do not develop pericytes. These lead to microvascular reductions and microaneurysms.
The platelet-derived growth factor is secreted by BMECs, and this leads to the recruitment
of pericytes during angiogenesis and, subsequently, vessel stabilization [
8
,
31
]. Cerebral
autoregulation might also be mediated by pericytes as they have been shown to express
receptors for angiotensin I, vasopressin, or endothelin-1 [
32
34
]. A recent review discusses
the functions of pericytes in the heart and brain, and shows there are still uncertainties up
to today [
29
]. Injured or degenerating pericytes have been reported in various studies of
neurological diseases, such as in Alzheimer’s disease (AD), Amyotrophic Lateral Sclerosis
(ALS), stroke, or mild dementia [12].
2.3. Astrocytes
Astrocytes received their name from their primarily star-shaped morphology. As the
most prevalent cell type within the CNS, these glial cells have more functions beyond
just providing support and structure. They are involved in synaptic formation, matura-
tion, plasticity, and neurotransmitter recycling [
35
,
36
]. Furthermore, their endfeet wrap
around almost the entire outer surface of the brain capillaries, and it is there that they
are in close interaction with BMECs. They secrete essential factors for BBB maintenance
and are involved in nutrient-waste exchange. Astrocytes enhance P-gp and glucose trans-
porter protein-1 expression in BMECs, as well as in metabolic enzymes [
37
]. Specialized
transporters and pumps regulate the CNS pH, fluid homeostasis, and electrical potential.
Prominent examples are the water channel aquaporin-4 (AQP-4) and the inwardly rectify-
ing K+ channel subunit 4.1 [
38
40
]. Astrocytes have been shown to harbor great anti- but
also pro-inflammatory potential, which is described in further detail in Section 3.1 [
41
44
].
2.4. Neurons
The understanding of the importance of neurons for BBB integrity has evolved over
the last few decades. They are not involved in BBB formation during early brain de-
Int. J. Mol. Sci. 2023,24, 12699 4 of 40
velopment [
45
]. Despite their approximate distance of 20
µ
m to the microvessels, they
communicate their need, via astrocytes, for additional oxygen or nutrients [
46
,
47
]. Astro-
cytes, in conjunction with pericytes, influence the vascular tone and blood supply in the
area to restore physiological conditions. Moreover, glutamatergic neurons can modulate
BBB integrity directly by increasing levels of glutamate, as well as influence the BBB efflux
transporter gene expressions and endothelial circadian genes [48].
2.5. Microglia
Microglia are often referred to as the immune cells of the brain. In a resting state,
microglia monitor the brain microenvironment, and they are always prepared to sense
antigens via their major histocompatibility complexes [
49
]. They have important functions
in both the adult and developing brain [
50
]. The phagocytic activity of microglia is crucial
for normal brain development by their removal of defective synapses and in helping with
synapse pruning [51]. Their role after injury or infection is described below in Section 3.1.
2.6. Oligodendrocytes
Oligodendrocytes belong to the macroglia and contribute as cells of the NVU to the
BBB. Oligodendrocyte precursor cells (OPCs) have crucial functions in BBB maintenance
and vessel formation throughout life. In return, they receive trophic support [
52
]. By
producing and maintaining the myelin sheath as an insulator around axons, they contribute
to the critical transmission of nerve impulses. In addition, OPCs are known to communicate
with other cells in the BBB, such as astrocytes and pericytes. For example, pericytes
stimulate OPCs via laminin-
α
2, which provokes the differentiation of OPCs into major
oligodendrocytes at the sites of damaged, demyelinated axons [53].
3. Main Players in Neuroinflammation
3.1. Cellular Components
Both brain resident cells, such as glial cells, and peripheral immune cells contribute
to neuroinflammation. Upon activation, microglia are polarized into a pro- or anti-
inflammatory phenotype depending on the stimulus [
51
]. The so-called M1 phenotype is
involved in the damage of surrounding neuronal and glial cells by secreting neurotoxic
factors, such as pro-inflammatory cytokines and chemokines like interleukin (IL)-6, tumor
necrosis factor-alpha (TNF-
α
), C-C motif ligand (CCL)-2, superoxide, and prostaglandin-
2 [
54
]. The M1 type is activated via the classical pathway by pro-inflammatory stimuli,
such as interferon-
γ
, the lipopolysaccharide (LPS) of gram-negative bacteria, or aggregated
pathogenic proteins (A
β
,
α
-synuclein and others) [
49
,
51
,
55
]. The M2 phenotype is involved
in tissue repair and wound healing by secreting anti-inflammatory mediators, such as
arginase-1 or chitinase-3. This phenotype can be induced by IL-4 or IL-13 in the alternative
pathway, or via acquired deactivation by IL-10 or the transforming growth factor-beta
(TGF-
β
). While both phenotypes are in homeostasis during acute stimulation, the pro-
inflammatory phenotype is predominant in chronic inflammation. Therefore, excessive
microglia activation can lead to a potentiating of tissue damage through a positive feedback
loop. This is the case in many neurodegenerative diseases, such as Parkinson’s Disease
(PD), AD, or ALS [56].
Similar to microglia, activated astrocytes also exhibit neuroprotective or neurotoxic
phenotypes. Liddelow et al. [
57
] proposed that, depending on the activation trigger, as-
trocytes develop different entities that are comparable to the M1/M2-type microglia. A1
astrocytes rapidly develop after acute CNS injury, such as CNS brain trauma or neuroin-
flammation. In response to the pro-inflammatory mediators that are secreted by M1-type
microglia, they induce a secondary inflammatory response [
49
]. This A1-type astrocyte
secretes neurotoxic factors that induce the rapid death of neurons and oligodendrocytes,
thereby driving neurodegeneration and disease progression [
57
]. Moreover, it sustains a
feedback loop that promotes further M1-type microglia, as well as leads to ECM and TJ
degradation via matrix metalloprotease (MMP) and vascular endothelial growth factor
Int. J. Mol. Sci. 2023,24, 12699 5 of 40
(VEGF)-A secretion [
58
]. In contrast, in ischemia A2, astrocytes have a neuroprotective
function, aiding neuronal survival and tissue repair [
57
]. The A1/A2 nomenclature, how-
ever, has been criticized by some researchers due to the high heterogeneity of astrocytes,
which extends beyond the binary nomenclature [5860].
Adhesion molecule expression is upregulated in BMECs during inflammatory condi-
tions so as to allow T cells to cross the BBB. P/E-selectins are involved in the initial step,
which is also called T cell rolling. The vascular cell adhesion molecule-1 and intercellular
adhesion molecule 1 lead to the arrest of CD4+ T helper (Th) cells, or other immune cells,
in the processes of capture, rolling, integrin activation, adhesion/arrest, crawling, and dia-
pedesis that occur across the BBB. In autoimmunity, it was shown that additional/added
melanoma or activated leukocyte cell adhesion molecules control T cell trafficking into
the CNS [
12
]. More detailed reviews of immune cell transmigration across the BBB can be
found elsewhere [6164].
3.2. Soluble Components
Cytokines mediate the communication within the immune system. Depending on the
specific cytokine and the receptor on the receiving immune cell, they can signal a particular
pathway for differentiation (such as Th1, Th2, or Th17). Chemokines are released by a
variety of cells, including endothelial cells. They play a role in embryonic development
and, depending on the chemokine, they recruit specific immune cells. The role of cytokines
and chemokines in neuroinflammation and the BBB has recently been reviewed, and the
reader is referred to these excellent review articles for more details [6568].
MMPs are a group of endopeptidases that are secreted by various cell types, including
members of the NVU [
69
]. Physiologically, they are important contributors in CNS develop-
ment, angio- and neurogenesis, as well as synaptic plasticity, learning, and memory [
69
,
70
].
Among others, MMP-2, -3, and -9 are engaged in BBB degradation and neuroinflamma-
tion [
70
,
71
]. They are activated through the proteolysis of the N-terminal pro-domain via
reactive oxygen species (ROS), IL-17, IL-1
β
, or TNF-
α
. In an active state, they degrade
ECM and TJ proteins with the subsequent activation of pro-angiogenic factors, such as
VEGF [
72
,
73
]. Additionally, they further promote a pro-inflammatory microenvironment
by cleaving more pro-MMPs and pro-forms of IL-1
β
or TNF-
α
[
71
]. Astrocytes, pericytes,
and microglia increase the expression of MMP-9 upon brain injury by increased ROS
through the albumin release or LPS [
71
,
74
,
75
]. Additional pro-inflammatory cytokines
are produced by the microglia that are activated via MMP-3 by apoptotic neurons [
76
].
The importance of different MMPs has been observed in a broad spectrum of pathological
BBB alterations, ranging from proteolysis after injuries, stroke-associated dementia and
hyperglycemia, multiple sclerosis (MS), over epilepsy, and many others [69,70,7780].
Lipid mediators are rapidly synthesized in multistep enzymatic pathways, and they
are bioactive lipids with functions in both health and disease. Depending on their structure,
they are classified into eicosanoids, lysophospholipids, and others. After synthesis, they
are released into the extracellular space, where they stimulate BBB cells and modulate their
functions via G-protein-coupled receptors. Through the activation of intracellular signaling
pathways, the paracellular permeability is ultimately increased [81].
ROS comprise a group of chemically reactive molecules that are involved with oxygen
(such as superoxide anions, hydrogen peroxide, and hydroxyl radicals), and they are
produced by several endo- and exogenous processes. They have the capability of oxidizing
or damaging biological molecules, and they have a dual role in the NVU [
82
84
]. Low
levels of ROS, together with reactive nitrogen species, act in signaling transduction for
normal cell function maintenance [
85
]. ROS also contribute to the regulation of blood
flow by vasodilation, the clearance of damaged cells, and tissue repair. Maintaining ROS
level balance underlies a complex interplay with cellular antioxidant systems, such as the
enzymes superoxide dismutase, catalase, and glutathione peroxidase [
86
]. Imbalances
in this regulation cause higher ROS levels, increased blood flow resistance, decreased
nitric oxide bioavailability, as well as increased apoptosis and immune response [
83
].
Int. J. Mol. Sci. 2023,24, 12699 6 of 40
ROS overproduction has also been associated with several neurodegenerative diseases,
especially AD. ROS are involved in several aspects of disease development, including BBB
breakdown (which interferes with brain energy supply and homeostasis),and increasing A
β
peptide deposition in vascular walls [
86
]. ROS promote a senescence-associated secretory
phenotype, which involves the secretion of several pro-inflammatory cytokines, MMPs,
and insoluble proteins/ECM [84].
4. In Vitro Modeling
Not every model is suitable for each research approach. Before starting to develop a
BBB in vitro model, the following questions should be considered:
What is the research question and the purpose of the experiment?
What basic biological requirements does the model need to fulfill?
What cells and what ECM are needed for the assays that are to be performed?
What time can be invested in each approach?
Does it need to be applicable for high-throughput screening?
Are there additional devices that are needed for the model?
How long are the cells cultured, and when is the endpoint?
An additional overview of the factors that need to be considered for in vitro modeling
is illustrated in Figure 1[87].
Figure 1.
Considerations for BBB/NVU model development. Adapted from Cameron et al. [
87
].
Created with BioRender.com.
4.1. Cell Types and Sources
The choice of cell types for each in vitro model is critical and depends on the re-
searcher’s needs and resources. Cells can be derived from animals (especially rodents,
but bovine and porcine tissues are also utilized) or humans. In general, they can be
Int. J. Mol. Sci. 2023,24, 12699 7 of 40
distinguished between immortalized cell lines, primary cells, and stem cells. Each cellular
model has certain pros and cons. Primary cells are isolated from postmortem brain tissue
(or from donor tissue that was collected during a surgical procedure), and they are then
frozen in a “low passage”. Thus, postmortem tissues are easier to acquire, but the material
isolated from surgeries often serves a better yield [
88
]. They exhibit characteristics close to
the in vivo NVU. However, they depend on a relatively high amount of cell-type-specific
growth factors, and they can only divide a limited number of passages before reaching
senescence, which affects cellular morphology and functionality [
89
]. Furthermore, isola-
tion of these cells raises ethical problems and requires high-level technical skills; moreover,
purchasing primary cells can be expensive [
90
]. Therefore, the immortalized cell lines of
primary cells were established to tackle this drawback. They possess most cell-type-specific
functions, are cost-effective, applicable for long-term culture, and can be expanded rapidly.
Unfortunately, immortalization can interfere with morphology and barrier formation. Thus,
different cell lines from the same cell type can have aberrant characteristics, which makes it
crucial to validate their features (such as cell-type-specific gene/protein expression and
barrier integrity before performing experiments [91,92]).
Over the last few years, an increasing number of in vitro BBB/NVU models have been
based on stem cells. The opportunity to differentiate them into various cell types with
healthy or pathological phenotypes has facilitated new perspectives for investigating the
NVU in health and disease [
93
]. There are several approaches for the use of stem cells as
neural stem cells have a low potential for self-renewal and show immune incompatibility
upon transplantation [
94
]. Therefore, they have been used to model diseases by inserting
certain gene mutations, and they are then re-injected into animal models rather than being
used for BBB models [
95
]. Mesenchymal stem cells are isolated from bone marrow. Due to
the secretion of growth factors that contribute to neuroregeneration and remyelination, they
have already been used to treat CNS diseases. Furthermore, since mesenchymal stem cells
express similar phenotypic markers as pericytes, they have successfully been implemented
in in vitro models as substitutes for primary pericytes [
96
98
]. However, they might be not
applicable for neuroinflammatory BBB in vitro models because they have been shown to
act immunosuppressively, and they might have AQP-4-regulating functions [
96
,
99
]. Finally,
induced pluripotent stem cells (iPSCs) are somatic cells that have been reprogrammed to a
state of pluripotency by utilizing an overexpression of specific transcription factors [
100
].
To achieve in-vivo-like paracellular barrier properties, iPSC-based human BBB models are
the models of choice; this is because they have a higher transendothelial electrical resistance
(TEER) than can be achieved in models utilizing other BMECs [
101
]. Furthermore, gene-
editing techniques, such as CRISPR/Cas9, enable for disease modeling through using
iPSCs [
102
]. Lippmann et al. [
103
] provided a detailed and widely used protocol for the
differentiation of iPSCs into induced brain microvascular endothelial cell-like cells (iBMEC)
for the purpose of BBB modeling. This protocol has been modified and further optimized
over the last decade [
104
106
]. However, attention needs to be paid to the cell identity,
as a recent paper revealed that these differentiation protocols produce a homogeneous
epithelial cell population instead of an endothelial one, despite exhibiting a high TEER.
Fortunately, this phenotype can be rescued by the overexpression of appropriate BMEC-
specific ETS transcription factors (such as ETV2,ERG, and FLI1). Although rescuing
the phenotype, Lu et al. [
107
] stated that, for reliable BBB-forming BMECs, more work
needs to be conducted, including thorough characterizations with the latest technologies.
Moreover, the narrow experimental window caused by de-differentiating iPSCs under
in vitro conditions results in high costs for culturing, and time-intensive procedures pose
limitations in iPSC models [
90
]. However, as an aspect of personalized medicine, patient-
derived iPSCs are becoming increasingly important. In addition to being used for modeling
diseases, these cells can also be helpful for the individual assessments of putative drug
responses [
93
,
100
,
102
]. For example, in monogenetic diseases, such as Huntington’s disease
(HD), the mutations are already present in patient cells and do not need to be created
additionally [108,109].
Int. J. Mol. Sci. 2023,24, 12699 8 of 40
Undoubtedly, BMECs are the most fundamental compartment of the BBB. Therefore,
culturing them as a monoculture can be utilized as a BBB model. However, as mentioned
above, other cell types crucially contribute to BBB integrity [
104
]. To date, due to their sim-
plicity, there are still monoculture models in use. These cultures are utilized in basic research
for toxicity and proliferation evaluations, transport experiments, and the characterization
of the secretion or immune response, especially when the response of only BMECs is aimed
at to be studied without the interference of other NVU cell types [
87
]. Additionally, it needs
to be noted that BMECs have a regional heterogeneity and an arterial-capillary-venous
zonation in vivo, which could lead to varying results in in vitro experiments depending on
the utilized BMEC source [
110
]. Very apparent differences in morphology and function are
visible in astrocytes and BMECs between gray and white matter, such as the expression of
specific receptors and transporters. The spatial density and orientation of the BMECs vary
depending on the location, and astrocytes show fibrous morphology with long processes in
the white matter while being more star-shaped in gray matter. Furthermore, oligodendro-
cytes are more abundant in white matter compared to the gray matter. Pericyte coverage
is also heterogeneous along the vessels. Differences have been reviewed in more detail
by Villabona-Rueda et al. [
111
]. Depending on the pathology that is to be modeled, these
differences need to be considered. For example, MS patients show different types of lesions
and BBB dysfunctions in their gray and white matter, and neurodegeneration in PD and
AD primarily affects gray matter [111,112].
4.2. Three-Dimensional ECM
In a 2D perspective, the ECM is set up solely by a layer of glial and/or neuronal cells
that are seeded onto a BM-mimicking substance. The currently used BM substitutes are
primarily collagen type I/IV and/or fibronectin [
87
]. To take ECM modeling to the third
dimension, a functional scaffold that supports appropriate cell growth and differentiation is
necessary. It must fulfill certain requirements, including ECM-comparable bio-physiological
properties (low stiffness, good hydrophilicity, elasticity, and degradation) and a low toxicity
for the cells [
113
]. Hydrogels are the medium of choice for these demands. They consist
of a biocompatible network of cross-linked polymer chains of natural or synthetic origin.
Natural polymers are either polysaccharide-based, such as hyaluronic acid or chitosan,
or protein-based, such as collagen, laminin, fibrin, or fibronectin [
114
]. Common substances
incorporate several proteins, such as gelatin, Geltrex
, or Matrigel
®
. The latter ones
have a thick structure and weak cross-linking, which makes them beneficial for 3D BBB
self-assembled models [
115
]. There are also plant-based materials available, such as,
for example, alginate, which is a polysaccharide that is derived from brown algae [
116
].
Although natural polymers reflect the physiological composition of the ECM, they have
several disadvantages, such as batch-to-batch variability, an (often) animal origin, and weak
mechanical properties, thereby leading to more rapid disintegration [
117
]. In contrast,
synthetic polymers are chemically defined and highly versatile, but they also poorly reflect
the physiological ECM composition. The primary synthetic substance that has been used
for BBB modeling is polyethylene glycol. Natural polymers combined with synthetic
polymers ensure a balance between mechanical strength and biocompatibility. These
hybrid hydrogels can have similarities with different CNS tissue characteristics, since ratios
between the constituent polymers, as well as functional groups or cross-linking agents,
can be modified. Consequently, their properties, such as degradation rate or stiffness, are
tunable, and they can be used to model healthy or pathological neural tissues. Examples of
often-used hydrogels are polyethylene glycol-hyaluronic acid, polyethylene glycol-collagen,
or gelatin-methacryloyl [
118
]. All hydrogels need some kind of cross-linking. This can
be achieved physically in the case of natural polymers, and this happens without the
addition of exogenous agents, which occur solely by the change in temperature or pH.
Hydrogels with synthetic compartments need chemical cross-linking, which is typically
photo-polymerization by an external substance that induces gelation [119].
Int. J. Mol. Sci. 2023,24, 12699 9 of 40
4.3. Transwell
With the first model being established in 1953, the transwell model setup has not
changed substantially since [
120
]. This approach is still widely used due to its cost-
effectiveness and high-throughput ability. In general, transwells consist of a polystyrene
multi-well carrier plate, which is available in several sizes, and its respective inserts.
The bottom of the inserts consists of a porous membrane that allows for the exchange
of nutrients and other molecules between the apical and basolateral compartments of
the insert. The pore size and density can vary depending on the application. Usually,
between 0.2 and 3
µ
m pores are used for transport studies and tissue engineering [
121
123
].
Thus, interactions between different cell types can be studied in co-culture models while
maintaining a physical separation. Smaller pore sizes are also recommended for the proper
formation of BMEC monolayers for barrier function assays [
121
]. When the chemotaxis
or invasion of cells is studied, membranes with larger pores of up to 12
µ
m are used [
124
].
The material of the membrane should be chosen individually depending on the application.
Polycarbonate membranes are made of a translucent thermoplastic polymer, and they are
employed in many cell culture studies due to their ease of use and low cost. However,
they are not fully transparent, and visualizing apically seeded cells under the microscope
requires cell labeling with fluorescent or chromogenic dyes. Furthermore, they have high
protein binding capacities, which is unsuitable for certain types of assays. Polyethylene
terephthalate membranes consist of polyester and exhibit low protein binding. They are
clear and offer good cell visibility. However, they have limited chemical resistance, which
is unfavorable for some applications. Polyvinylidene fluoride membranes have uniform
or asymmetric pores, low protein binding, as well as high chemical and thermal stability.
They are suitable for a wide range of applications, but are also more expensive. Coating the
membranes with BM components, such as collagen type IV or fibronectin, is recommended
for better cell attachments and morphologies [
123
]. Recent models used the co-cultures of
human immortalized cell lines [
123
,
125
], primary cells [
122
], or iPSCs [
126
129
]. Some of
them incorporated hydrogels on one side of the insert with glial cells to obtain a more phys-
iologically relevant ECM setup [
129
131
]. Transwell co-cultures are often criticized for their
lack of physiological relevance due to the absence of contact between different cell types.
The thickness of the membrane, which is typically at least 10
µ
m thick, restricts the cell–cell
interactions across the membrane. This problem was addressed by Zakharova et al. [
132
],
who developed polydimethylsiloxane (PDMS) membranes with tunable thicknesses and
pore sizes. They fabricated an optically transparent, 2
µ
m thin membrane that enabled for
a lower permeability in BMECs when co-cultured with astrocytes when compared to other
membrane substances.
4.4. 3D Models
Three-dimensional models have made some advances compared to the classical tran-
swell models. First, a 3D ECM provides a more in-vivo-like environment than the planar
cell layers of most transwell models. Second, aside from membrane-based microfluidic
models, the different cell types are in direct contact with each other, which facilitates the
interaction and exchange of secreted growth factors or other regulatory molecules. Third,
a medium flow mimicking the shear stress of the blood circulation can be implemented.
The neuroinflammatory response profile to TNF-
α
stimulations in 2D vs. 3D models
was investigated by Herland et al. [
133
] in mono- and co-culture models by utilizing tran-
swell and microfluidic BBB-on-a-chip setups. Co-culturing primary BMECs with astrocytes
and pericytes under flow conditions resulted in significantly higher cytokine secretion (i.e.,
of IL-6, the granulocyte colony-stimulating factor) in the 3D model. Cucullo et al. [
134
]
showed, in 2011, the importance of shear stress on BMECs, as it altered the gene expression
patterns of junctional proteins, CYP450 proteins, ion channels, drug transporters, adhesion
molecules, and integrins. Notably, evidence has arisen of the notion that shear stress
does not affect cell morphology, but rather it tightens the barrier itself by an increase in
adherens junction and TJ protein expression [
87
]. The diameter of the vessel (channel,
Int. J. Mol. Sci. 2023,24, 12699 10 of 40
tubing), the dynamic viscosity of the fluid (medium), as well as the flow rate need to be
adjusted to get close to the physiological amount of shear stress that is needed in the model.
The in vivo shear stress is approximately 0.3–2 Pa (4 to 30 dyn/cm
2
), and this is with a
dynamic viscosity, which is dependent on the rheology of blood and is usually calculated
as 3.5–5.5~cP [
87
,
135
,
136
]. The calculation of the shear stress adjustment is described in
detail elsewhere [
87
,
137
,
138
]. For medium-flow generation in microfluidic devices, syringe
and peristaltic pumps are widely used in several other methods [
139
]. However, there
are new advances in pumping systems that passively pump media in a non-mechanical
manner [
140
]. It is important to note that this is a continuous rather than pulsatile flow, as
is keeping the shear stress on a moderate, physiological level [
141
]. Faley et al. [
142
] created
a 3D microfluidic model with a tubular structure filled with the monolayers of endothelial
cells. The authors compared different endothelial cell lines, ranging from human umbilical
vein endothelial cells to human dermal microvascular endothelial cells, as well as to two
iPSC-derived BMEC lines. The iBMECs showed a 10–100 times lower permeability to
different-sized markers compared to the other cell lines. Moreover, the model maintained
the barrier function and efflux transporter activity for up to 21 days under perfusion condi-
tions. It is worth mentioning that Faley et al. saw the best results with a subphysiological
wall shear of 0.3 dyn/cm
2
. Elevated flow rates led to an increase in permeability and
angiogenic sprouting.
One common problem regarding microfluidics is the formation of bubbles, which are
generated when plugging or removing pipes, exchanging reagents, or if the systems are
not fully tight. These bubbles tend to accumulate at right angles in the microfluidic chips,
particularly at the device inlets. Moreover, these bubbles cause blockages in the channels
known as “dead zones.” As such, they lead to flow interference, which increases the
pressure and causes cell membrane damage. To address this problem, active (for example
lasers or acoustic generators) and passive (which do not require additional equipment)
de-bubblers can be used [143].
4.4.1. Microfluidic Models
The first microfluidic models were described in the late 1990s, and they consisted
of a hollow fiber cartridge system that was composed of a bundle of porous polypropy-
lene fibers. Through the fibers, there was a medium flow applied by a variable-speed
pulsatile pump. Endothelial and glial cells could be seeded through separate loading
ports. The tubing was made of gas-permeable silicon [
144
,
145
]. Continuing up to the
present day, this original setup has been further developed and optimized, and the research
on novel microfluidic devices is still intensive. However, the major components are still
present: (1) a carrier in which the cells are incorporated, (2) an ECM compartment, (3) a
BMEC monolayer, and (4) medium flow. The material of microfluidic chips has mainly
been PDMS since it fulfills most needs. It is biocompatible, easy to fabricate, cost-efficient,
and transparent. However, one limitation of the material is its tendency in the absorption
of hydrophobic molecules. This problem is minimized by conducting protein modifications
at its surface [146].
However, there are versatile possibilities for the chip appearance, and there are four
major designs [
137
]. The sandwich design comprises a PDMS chip with two superposed
channels, which are separated by a polycarbonate or PDMS membrane. Glial cells or
neurons are seeded into the upper channel, which mimics the CNS matrix. The lower
channel reflects the capillary with the BMECs being seeded onto the membrane. Similar to
this setting is the parallel design. Cells are seeded not in vertically but in horizontally sepa-
rated channels. These channels have small interconnecting pipes with a diameter of 3
µ
m
to enable communication between the cell types [
147
]. As a further development of the
parallel design, chips with two channels for BMECs and an additional channel in between
with a CNS matrix replicate have been established. This design is membrane-based as well,
but incorporates a hydrogel-aided compartment for the ECM [
148
]. The communication
between cells can be further enhanced by 3D tubular structure designs. Without the need
Int. J. Mol. Sci. 2023,24, 12699 11 of 40
for a membrane, the cells are in direct proximity. Cylindrical channels lined with BMECs
are surrounded by a cell-containing hydrogel, thus reflecting the ECM. The microvessel
structure can either be achieved artificially or with self-assembled microvessels in hydro-
gels [
149
]. Campisi et al. [
150
] combined an iBMEC network with human brain pericytes
and astrocytes within a single fibrin hydrogel. After vasculogenesis, the model exhibited
physiologically relevant structures and functionality. Organs-on-a-chip microfluidic models
in various setups can be purchased as well. Different companies have launched chips for
all kinds of research areas. The reviews of Nikolakopoulou and Jagtiani [
151
,
152
] sum up
the currently available products.
4.4.2. 3D Bioprinting
Over the last few decades, 3D bioprinting has pushed the precision of BBB model
fabrication to a new level. Templates are designed by computer-aided design software and
then printed by specialized 3D printers. Depending on the printing technology, precise
structures with a resolution down to 100
µ
m (extrusion-based bioprinting) and even 10
µ
m
(laser-assisted bioprinting) are possible [
113
]. There are different fabrication techniques,
each of which imply strengths but also limitations [
119
]. Inkjet bioprinting offers a high
speed and resolution with low cost. It is favorable for printing cells as it ensures a high cell
viability of up to 90% [
153
]. A printer head connected to a cartilage creates droplets of bioink
by a thermal or piezoelectric actuator [
154
]. To prevent the printer head from clogging,
printing high cell densities or viscous material need to be avoided [
119
]. Microextrusion
bioprinting allows for the printing of high-viscous material in a continuous strand through
a syringe via a screw plunger or through air pressure. Compared to other bioprinting
methods, it is relatively cheap and simple [
154
]. However, it compromises cell viability
by exposing them to high mechanical stress. Therefore, extrusion-based bioprinting is
used more for scaffold and sacrificial structure printing [
155
]. In laser-assisted bioprinting,
the bioink is suspended at the bottom of a donor layer, which has a ribbon structure on
top. The printing of droplets is achieved by an absorbing layer that is stimulated by a
laser pulse. This pulse leads to an evaporation of the donor layer, creating a bubble at
the interface of the bioink layer, which is pushed onto the substrate. There, the bioink
droplet is then cross-linked. This allows for the printing of very precise structures with a
broad range of bioink viscosity. Due to the contact-free printing, cells do not experience
mechanical stress. Although cell viability is high, the technique’s high costs restrict its use in
research [
119
,
154
]. Different kinds of light-based bioprinting (stereo-, soft-, and two-photon
lithography) are rapid and highly precise methods, and they are conducted by a hydrogel
layer being exposed to a light pattern (usually UV or near-UV wavelengths); through this,
the illuminated area is cross-linked. This process is repeated vertically until the whole
construct is built. Scaffolds can be printed in such precision that Mariano et al. [
156
], for the
first time, developed a 1:1 scale biomimetic BBB model. Two years later, the same group
published a model with even thinner capillary walls of only 2
µ
m instead of 10
µ
m, as
was the case in the prior model [
157
]. Despite the high cell viability and fast fabrication,
a drawback of these methods is the limitation to single-material structures [119,154].
The printable material in 3D bioprinting is highly diverse. It needs to be distinguished
between substances that incorporate cells and those that are utilized for frameworks or
for dissolving support. Bioinks with cells are hydrogels, as described in Section 4.2. The
frameworks for the chips themselves or scaffolds, where cell-laden hydrogels are added in
between, are usually made by highly viscous substances that retain their structure. Resins
or PDMS are the materials of choice [
157
,
158
]. Sacrificial inks can be used to print hollow
structures into matrices since they dissolve under specific circumstances, like when there
are temperature changes. Examples are Pluronic (F-127) or polyvinyl alcohol [
158
]. Even the
chips themselves can be printed, which is achieved by utilizing transparent polymers like
transparent poly methyl methacrylate) or duralumin [
158
]. Wang et al. [
159
] developed an
Objet VeroClear photopolymer-based chip with a perylene-C coating to ensure transparent
appearances, chemical resistance, and biocompatibility.
Int. J. Mol. Sci. 2023,24, 12699 12 of 40
4.4.3. Cell Aggregates (Spheroids and Organoids)
Spheroids and organoids are both self-assembled cell aggregates in a static three-
dimensional environment. However, they differ in composition, complexity, and size.
Spheroids are more simple cell structures that are primarily generated by cell lines, primary
cells, or tumor cells/tissues in suspension or non-adherent conditions. They do not require
an ECM, and they usually do not exhibit a defined tissue architecture. Organoids are
derived from culturing iPSCs, embyronic, or neural stem cells, and they can recapitulate
many aspects of tissue and organ function through which to assess neurotoxicity and to
develop new drugs. However, cell differentiation requires an ECM and a growth factor
cocktail [
160
]. The first approaches were established a decade ago and are progressing
rapidly [
161
,
162
]. For example, Nzou et al. [
163
] studied the suitability of their human
cortex organoid model for drug development by assessing the BBB impairment under
hypoxic and neuroinflammatory conditions, as well as the impact of ROS and inflammation-
reducing reagents. One major advantage of organoids is the possibility of modeling
different brain regions. Thus, organoids resembling the human midbrain can be developed,
including astrocytes, oligodendrocytes, and dopaminergic neurons that are myelinated and
exhibit synaptic connections [
162
]. Furthermore, the self-assembled structures can be set up
based on patient iPSCs, which allows for disease- and patient-specific modeling. In using
this approach, the potential neurotoxicity for the patient can be screened and might also be
helpful in therapy development in the future [
163
]. A disadvantage is that organoids are
limited in size (around one millimeter) due to the lack of nutrient availability in their core.
This often leads to cellular stress within the CNS-mimicking organoid center, resulting in
necrotic cells with fragmented nuclei. This issue could be solved by vascularization within
the organoid. However, this is still a major challenge for researchers in this field [
164
]. A
study by Wörsdörfer et al. [
165
] suggested the implementation of mesodermal progenitor
cells into organoid models in order to tackle the lack of vessels within the 3D structure. Their
tumor organoid model incorporated blood vessels, as well as BM and cell–cell junctions.
Additionally, the mesodermal progenitor cells also differentiated into microglia-like cells
that could potentially be used for tumor modeling but also neuroinflammation in general.
Ham et al. [
166
] developed vascularized cerebral organoids by adding VEGF from the
beginning of the differentiation of embryonic stem cells, thus creating open-circle vascular
tubes with a two-layer structure and BBB characteristics without disturbing neurogenesis.
However, in long-term cultures, the density of blood vessels decreased and four-month
old organoids had distorted open-circle morphologies, which was possibly caused by the
missing blood pressure. Therefore, the importance of shear stress within, as well as on, the
outer surface of cerebral organoids is still a matter of research [89,166].
4.5. Model Validation
4.5.1. Cell Identity and Viability
The validation of the cell-type identity is not only for the iPSCs of special impor-
tance. Although manufacturers promise the expression of their commercially immortalized
and primary cell lines, they still might change their functions and morphology due to
immortalization or cell culture conditions, as well as due to the number of passages. There-
fore, cell-type-specific markers should be assessed via immunofluorescence, Western blot,
qPCR, or flow cytometry before the performing experiments [
119
]. BMEC identity is
usually assessed by a combination of general endothelial cell markers—such as vascular
endothelial-cadherin (CD31), specific markers for functionality, and TJ protein expres-
sion—which is a topic that will be covered in the next section. Astrocyte markers are
primarily GFAP and AQP-4. While GFAP can also be found on other glial cells in the devel-
opment of natural tissues, it is expressed only by astrocytes in a differentiated neuronal
culture [
167
]. AQP-4 is most abundantly expressed by astrocytic endfeet, but other ependy-
mal cells show low expressions as well [
168
]. Both markers can be instrumental in disease
modeling as their expression enhances or decreases in the course of pathogenic changes
or through the reactivity of astrocytes [
167
]. Pericyte identification can be challenging if
Int. J. Mol. Sci. 2023,24, 12699 13 of 40
they are isolated from tissues with vascular smooth muscle cells, but this can be neglected
in the microvasculature, as this cell type is not usually found there. The platelet-derived
growth factor receptor-
β
and neural/glial antigen-2 are accepted markers for pericyte iden-
tity [
169
]. The oligodendrocyte mature markers are oligodendrocyte transcription factors 1
and 2, or myelin-associated proteins, including myelin basic protein, 2
0
,3
0
-cyclic nucleotide-
3
0
-phosphodiesterase, as well as myelin oligodendrocyte glycoprotein (MOG) [
119
,
170
].
The latter is of special importance for the investigation of inflammatory demyelinating
diseases. Depending on their reactive state, specific microglial markers can be used. Resting
microglia express CD11b and CD45. When entering a pro-inflammatory phenotype through
external stimuli, M1 microglia express CD86 and inducible nitric oxide synthase, whereas
anti-inflammatory M2 microglia upregulate their arginase-1 and CD206 expression [
171
].
Furthermore, the development, maturation, and functionality of neural cells can be investi-
gated in 3D fabrications. For neural differentiation, nestin is often used as a progenitor and
as a neural stem cell marker [
172
]. The neuron-specific class III
β
-tubulin and its antibody
TUJ are markers for premature or undifferentiated neural cells. Major neurons can be
identified by microtubule-associated protein 2. Mature neuronal subtypes can be further
divided by the markers gamma-aminobutyric acid (GABA) and its enzymes GAD65/67 for
interneurons; tyrosine hydroxylase for dopaminergic neurons; tryptophan hydroxylase for
serotonergic neurons; and choline acetyltransferase for cholinergic neurons [119].
Assessing the cellular viability is particularly important in 3D models since nutrient
and oxygen availability is not always ensured; moreover, it is dependent on the thickness
and porosity of the matrix and the chip itself. Furthermore, extensive UV cross-linking
during fabrication is compromised with cell viability. Besides checking cell morphology
and the counting tools that are available for confocal microscopes from 3D stacked image
sequences (such as ImageJ plugin 3D Objects Counter), several reagents can be used for
viability evaluation [
173
]. Tetrazolium salts (like MTT) are reduced by living cells into
formazan dyes, which can then be quantified via visible light absorbance measurements.
Likewise, resazurin-based substances, such as Alamar Blue, can be used for estimating the
living cell population, whereby the number of living cells is estimated through reduction
reactions that produce the red fluorescent resorufin. The advantage of tetrazolium- and
resazurin-based chemicals is their simple application to the culture medium, which is
incubated for a defined time and can be measured immediately afterward. However, these
methods do not reflect the dead-cell compartments, and reductions in a 3D ECM occur only
slowly [
174
]. For the evaluation of both viable and dead cells, a mixture of Calcein-AM
(which can easily enter living cells) and ethidium homodimers (which only enter dead
or damaged cells) is the combination of choice, if it is applied with different fluorescent
tags [
119
]. Apoptosis can be evaluated in assays with markers like caspase-3 and -9, as well
as Bcl-2 and Bax, or with terminal deoxynucleotidyl transferase dUTP nick end labeling
assays. Since apoptosis is detectable in diverse pathological conditions, such as traumatic
brain injury (TBI) or ALS, assays should be also included in the in vitro modeling of these
diseases [175,176].
4.5.2. TJ Formation and Permeability
BBB permeability is most frequently tested by TEER measurements. The presuppo-
sition is that the barrier has a setup that allows for measurements on the basolateral and
apical side. As such, an electrical field is created at both sides of the barrier, and the resis-
tance is measured by electrodes. The higher the resistance, the stronger the barrier function
and the more tightly the cell junctions are connected. The advantage of this method is the
non-invasive and rapid manner in which the barrier can be assessed.
TEER =RA
where Rdepicts the electrical resistance in
(which is equal to the measured resistance
minus the resistance of the blank) and Adescribes the growth area in cm
2
. The average
physiological electrical resistance in vivo is around 1800
cm
2
, but this can reach values as
Int. J. Mol. Sci. 2023,24, 12699 14 of 40
high as 8000
cm
2
[
177
,
178
]. However, only very few current in vitro platforms can reach
values as high as this [
159
,
179
,
180
]. Although TEER measurement is the gold standard
for permeability assessments in transwell and microfluidic model systems, the technique
lacks consistency in the results. A critical review from Vigh et al. [
181
] highlights the
range of values that depend on the device used for measurements. The authors stated
that the correct interpretation of values, as well as the comparison between the models,
is only possible with an explicit description of the technical parameters and the setup.
Palma-Florez et al. [
182
] tried to address this limitation by developing a BBB microfluidic
chip with an integrated micro-TEER device that was in a close proximity to the barrier. To
improve TEER, certain substances can be added. For instance, synthetic glucocorticoids
like dexamethasone or hydrocortisone (cortisol) have been found to increase TEER values
in vitro. However, they may impact experimental results, particularly in studies that are
related to inflammation as glucocorticoids have anti-inflammatory properties [183,184].
Permeability can also be assessed by specific markers that are non-toxic and do not
bind, or through those that become internalized by the cells within the model. Furthermore,
they should be metabolically inert, available in different molecular sizes, quantifiable,
and reliable [
12
]. Detailed descriptions of how permeability needs to be calculated in 2D
and 3D setups were published by, for example, Hajal et al. [
185
] and Wong et al. [
186
].
Kadry et al. [
12
] provided an overview of the different reagents that can be used to test
permeability within the context of various approaches.Since none of the substances that
are currently available fulfill all the aforementioned requirements, the choice of marker
needs to be made based on the specific research question. Both small (for examining small
changes) and large molecules (to assess BBB integrity loss) can be used as markers for
BBB dysfunction. Since none of the substances currently available fulfill all the necessary
requirements, the choice of marker for BBB dysfunction must be made based on the specific
research question. Small molecules are suitable for studying small changes, while large
molecules can be utilized to evaluate BBB integrity loss. As such, a combination of different
markers is the most reliable approach through which to evaluate BBB integrity
[12,187].
TJ
proteins can either be directly evaluated via immunofluorescence, qPCR, or via Western
blotting. The most commonly accepted markers for the BBB are claudin-5, ZO-1, and oc-
cludin. Additionally, the markers for efflux transporters, such as P-gp, BCRP, Mrp, or solute
carriers can be utilized to ensure BMEC functionality and polarity [
187
]. A recent study
by Nakayama-Kitamura et al. [
123
] set up evaluation parameters for the in vitro human
BBB likeness that is needed for drug development. The parameters comprised TJ markers
(claudin-5 and ZO-1), TEER, the endothelial cell marker CD31, transporters (P-gp, glucose
transporter protein-1, and BCRP), and receptor-mediated transcytosis (transferrin receptor).
Additionally, they checked the permeability of caffeine (which is usually BBB-permeable)
and Lucifer Yellow (which is impermeable if the BBB is functional).
5. Disease-Specific Modeling of Neuroinflammation
The following section will discuss the various conditions that involve neuroinflamma-
tion. However, since there are more than 600 neurological diseases known today, it will
only cover a fraction of the most prominent examples that come with BBB impairment [
188
].
Key features of the discussed neurological diseases are summarized in Table 1.
Int. J. Mol. Sci. 2023,24, 12699 15 of 40
Table 1. Key features of the neurological diseases with BBB impairment.
Disease Key Features References
CNS Autoimmune Diseases
Complex interplay of genetic and environmental factors
Disease onset and severity are driven by triggers, such as infections or tumors
Autoantibodies and reactive T cells
[189,190]
Multiple Sclerosis
Chronic inflammatory demyelinating disease with autoimmune aspects. Th1 and Th17 CD4+ T cells might be autoreactive,
thus producing interferon-
γ
(by Th1 cells), IL-17, and IL-22 (by Th17 cells), which dysregulate TJ proteins and upregulate
adhesion molecules
Human leukocyte antigen gene polymorphisms and viral infections (Epstein-Barr Virus)
Inflammatory processes are provoked by neurotoxic astrocytes and microglia, which attract immune cells and lead to
progressive or relapsing axonal demyelination with oligodendrocyte apoptosis, astrogliosis, axonal loss, and secondary
neurodegeneration
[16,191193]
NMDARE NMDA receptor autoantibodies lead to an internalization of the receptor
Autoimmune Encephalitis with psychiatric symptoms, autonomic fluctuations, and seizures [190,194]
NMOSD
The majority of patients are seropositive for the antibodies that are against AQP-4 on the endfeet of astrocytes that induce
complement-dependent cytotoxicity
Association with acute optic neuritis, myelitis, area postrema syndrome, and distinct MRI lesions
[189,195,196]
MOGAD
Patients have autoantibodies against MOG on the oligodendrocytes that can activate FcR-mediated antibody-dependent
cellular cytotoxicity and which complement-dependent cytotoxicity
Association with acute disseminated encephalomyelitis, optic neuritis, transverse myelitis, brain stem syndrome, and encephalitis
[189,197]
CNS Infections
Bacterial Infections
The endotoxin LPS of gram-negative and lipoteichoic acid from gram-positive bacteria leads to BBB disruption via toll-like
receptor-4, thereby inducing pro-inflammatory cytokines for the innate immune response
The endotoxin hypothesis suggests that gut-bacterial endotoxin causes/contributes to neurodegeneration
[198,199]
Viral Infections
Entry via transcytosis, and the infection of BMECs or CNS-invading immune cells occurs as a Trojan horse
HIV-1, SARS-CoV-2, and others reduce TJs and upregulate adhesion molecules/cytokines
Can cause/trigger other neurological disorders, such as dementia, epilepsy, autoimmune diseases, etc.
[200,201]
Int. J. Mol. Sci. 2023,24, 12699 16 of 40
Table 1. Cont.
Disease Key Features References
Acute CNS Injuries
Stroke
Ischemia (in 85% of cases) with nutrient and oxygen deprivation in the hyperacute phase (<6 h), whereby the TJ opening
occurs due to hypoxia-induced ROS and transcription factors (hypoxia-inducible factor-1
α
). This is followed by cytotoxic
(ionic) edema due to the cell swelling that occurs because of ion and water homeostasis imbalance
A refractory period with increasing BBB openings (hours or days) leads to plasma protein entry in the CNS, possibly resulting
in a vasogenic edema by fluid increase due to BBB openings and the extravasation of plasma proteins
Potential hemorrhagic conversion with a loss of microvessel integrity and the entry of blood components into the CNS
[47,81,202204]
Traumatic Brain Injury
Inflammation by microglia and astrocyte activation (via albumin/DAMPs) and by neutrophilic infiltration; chronic inflamma-
tion might be linked to enhanced white matter-/neurodegeneration and encephalopathy
Elevated neuronal stimulation and toxicity through the overproduction of glutamate receptors and GABA-receptor internal-
ization
Further BBB breakdown and dysregulated metabolism by ROS and other pro-inflammatory molecules, as well as in the
sub-acute stages of TBI that is enhanced by NOX and activated by MMPs
[75,203,205,206]
Epilepsy
Recurrent seizures whose frequency has been shown to correlate positively with BBB breakdown
Higher K+ influx to the CNS due to leukocyte–BMEC interactions. Astrocytes and microglia produce pro-inflammatory
mediators that lower TJ protein expression and BBB tightness. The extravasation of serum proteins and immune cells augment
the inflammatory responses
The downregulation of the K+ channel subunit 4.1 (via albumin binding to TGF-
β
receptor) and excitatory amino acid
transporter-2 (via TNF-α) in astrocytes cause an increase in the K+ and glutamate that enhance neuronal hyperactivity
[203,207]
CNS Tumors
Primary Tumors Glioblastoma multiforme (from astrocytes) is the most aggressive one, but does not metastasize
Cancer cells release exosomes with VEGF-A and pro-inflammatory cytokines, which lead to a TJ decrease [12,208]
Brain Metastases A total of 30% of brain cancers, often derived from lung cancers, breast cancers, or melanomas
Microvessels have heterogeneous permeability with their active efflux of molecules, enhanced angiogenesis, and inflammation
[152,203]
Int. J. Mol. Sci. 2023,24, 12699 17 of 40
Table 1. Cont.
Disease Key Features References
Neurodegeneration
Alzheimer’s Disease
The most common cause of progressive cognitive impairment worldwide (increasing with the rising percentage of aged
people) that has genetic and environmental risk factors
Genetic forms (in total 1%) are caused by mutations in the genes APP,PSEN1, or PSEN2
Neuroinflammation and degeneration, Aβplaques, and neurofibrillary tangles (hyperphosphorylated tau)
[16,209]
Amyotrophic Lateral Sclerosis
Muscle weakness and atrophy with a progressive decrease in motor functions through motor neuron degeneration
The 43 insoluble TAR-DNA-binding protein aggregations in the soma of motor neurons exhibit neurotoxic properties. This
also occurs in other misfolded protein aggregates (with mutations in SOD1 or C9orf72) in familial forms (10%)
Glutamate-mediated excitotoxicity, increased oxidative stress, mitochondrial dysfunction, and neuroinflammation
[210212]
Parkinson’s Disease
Age-associated neurodegenerative disease with dopaminergic neuronal loss in the substantia nigra
Neuroinflammation mediated by Lewy body depositions (the aggregations of α-synuclein oligomers or fibrils)
Genetic factors, environmental toxins, and mitochondrial dysfunction play a role
[213216]
Huntington’s Disease
Glutamine (CAG) repeat expansion mutations in the HTT gene
Affected areas are the basal ganglia (i.e., the dorsal striatum and cortical compartments)
Reactive microglia and astrocytes play major roles in neuroinflammation generation
[108,217219]
Abbreviations:
A
β
, amyloid-beta; BBB, blood–brain barrier; CNS, central nervous system; DAMPs, damage-associated molecular patterns; GABA, gamma-aminobutyric acid; HIV-1,
human immunodeficiency virus-1; IL, interleukin; LPS, lipopolysaccharide; MOG, myelin oligodendrocyte glycoprotein; MOGAD, MOG-associated antibody disease; NMDARE,
N-methyl-D-aspartate receptor encephalitis; NMOSD, neuromyelitis optica spectrum disorders; NOX, nicotinamide adenine dinucleotide phosphate oxidase; ROS, reactive oxygen
species; TGF-
β
, transforming growth factor-beta; Th, T helper; TJ, tight junction; TNF-
α
, tumor necrosis factor-alpha; SARS-CoV-2, severe acute respiratory syndrome coronavirus type 2;
and VEGF-A, vascular endothelial growth factor-A.
Int. J. Mol. Sci. 2023,24, 12699 18 of 40
5.1. Neurological Autoimmune Diseases
BBB dysfunction is an early event in MS, which is followed by immune cell infiltration
and the extravasation of plasma proteins. How the BBB contributes to MS pathogenesis and
progression is discussed elsewhere [
46
,
220
223
]. Nishihara et al. [
224
] showed that different
Th subsets had comparable migration rates through the BBB in a transwell setup under
inflammatory conditions (mediated by TNF-
α
and interferon-
γ
). However, it was observed
that the Th1 cells preferentially migrated through the barrier in non-inflammatory condi-
tions. The authors also investigated the movement through the blood-cerebrospinal fluid
barrier and saw a 10- to 20-fold higher migration compared to the BBB model, especially
regarding Th17 cells. Based on these findings different Th cell subsets might use different
anatomical routes to enter the CNS. The Wnt/
β
-catenin pathway is involved in numerous
crucial regulations of embryonic development and adult tissue homeostasis. Among other
functions, it is important for maintaining the BBB’s integrity and restoring it in disease,
including in the case of neurodegenerative diseases like AD, PD, and ALS [
225
,
226
]. In MS,
in vivo and in vitro studies have revealed that abnormalities in Wnt/
β
-catenin signaling
contribute to re-myelination failure. OPCs cannot detach from the vasculature and are
unable to migrate to the demyelinated region. Furthermore, un-detached OPCs lead to
perivascular clustering and the secretion of Wif1, which reduces Wnt and TJ integrity [
227
].
Derada Troletti et al. [
228
] showed that endothelial to mesenchymal transition, which
they provoked by TGF-
β
and IL-1
β
additions, mediated the inflammation-induced BMEC
dysfunction, and they found that this might also play a role in MS pathophysiology. The
main transcription factor of endothelial de-differentiation, ETS1, has also been shown to be
associated with BBB breakdown [
229
]. Cerutti et al. [
230
] developed an in vitro technique
for studying the interaction between BMECs with human leukocytes in a microfluidic
model, which was coupled with live-cell imaging that could also be advantageous for
leukocyte extravasation into the CNS in MS in vitro studies. It appears to be impossible
to create a universal in vitro model for neuroinflammation in MS due to the reliance of
adaptive and innate immune cells on the disease’s development, as well as due to the
specific region of the brain that is affected [192].
For drug development, BBB in vitro models are urgently needed to investigate the
effects of treatments for retained integrity or barrier restoration. In the CNS, the differ-
ent pathogenic autoantibodies have different modes of action. Without proper treatment,
CNS autoimmunity can have severe effects in patients, leading to cognitive impairment,
seizures, or even comas in NMDA-receptor encephalitis (NMDARE), or MS-like relapses
in AQP-4 antibody-positive neuromyelitis optica spectrum disorders (NMOSD). Acute
relapses are typically treated with general immunosuppression via the use of corticos-
teroids. Although corticosteroids decrease the permeability of the BBB and the produc-
tion of pro-inflammatory cytokines, chemokines, and cell adhesion molecules, they can
also have severe side effects. Therefore, more targeted therapeutics are needed [
231
].
Takeshita et al. [
232
] developed a tri-culture transwell model, whereby they applied the
antibodies from NMOSD patients who were with or without the IL-6 receptor blocking
therapeutic antibody satralizumab. The addition of satralizumab led to higher TEER val-
ues and decreased the transmigration of peripheral blood mononuclear cells through the
barrier. When studying the recombinant antibodies from NMOSD patients’ cerebrospinal
fluid, Shimizu et al. [
233
] observed a strong binding in one of their antibodies against
glucose-regulated protein 78, which is an endoplasmic reticulum chaperone that has also
been found on the surface of BMECs in vitro. Therefore, the hypothesis that arose was that
other non-disease-defining autoantibody targets are relevant for the BBB breakdown in
these diseases, and that they could either be possible therapeutic targets or even be em-
ployed for the enhanced accessibility of immunotherapeutics into the brain. To investigate
this hypothesis, Li et al. [
194
] analyzed the human monoclonal antibodies from patients
with NMDARE and GABA
B
-receptor encephalitis for non-disease-defining antibody target
identification. Antibodies that showed vascular binding in mouse brain sections were
applied to a BMEC monoculture transwell system to evaluate if there were effects on
Int. J. Mol. Sci. 2023,24, 12699 19 of 40
their microvasculature.One of their monoclonal antibodies decreased TEER and occludin
expression, which was also confirmed in vivo. Finally, myosin-X was identified as a novel
target epitope.
5.2. CNS Infections
Many bacterial species are able to infect the CNS with intact, but also especially im-
paired, BBBs via diverse mechanisms [
234
]. Brown et al. [
235
] investigated, in their 3D
microfluidic NVU dual chamber model, the impact of inflammatory stimulation on the
BBB, and this was induced by LPS (mimicking a systemic bacterial infection) or via pro-
inflammatory cytokines (which could result in a local or systemic inflammation). By apply-
ing either 100 µg/mL of LPS or a 100 ng/mL cytokine mix (of IL-1, TNF-α, and monocyte
chemoattractant protein-1,2) to the vascular compartment, inflammation was induced,
and the metabolic profile on both sides was investigated over 24 h of exposure. An activa-
tion of pro-inflammatory cytokines was found in both vascular and brain compartments,
whereas, at later time points, this activation was only seen in a subset in both compartments.
This could suggest the activation of anti-inflammatory cytokines at later time points within
the brain section, whereas the vascular compartment remained as more pro-inflammatory.
The bacterial infection of the CNS that causes bacterial meningitis has been modeled with
different pathogens, and this was achieved with several BMEC cell lines in the transwell
models [
236
]. Streptococcus pneumoniae (among others) was shown to upregulate VEGF via
hypoxia-inducible factor-1αinduction.
Consequently, BBB permeability was elevated, thus enabling more bacterial transmi-
grations into the CNS [
237
]. VEGF secretion was also observed by Caporarello et al. [
238
]
when infecting their BMEC-pericyte co-culture BBB model with Haemophilus influenzae type
a. The authors demonstrated that different adenosine receptors on both cell types were
activated upon infection and that VEGF, was released by the cells, which can cause pericyte
detachment, BMEC proliferation, and BBB breakdown. More recently, the impact of the gut
microbiome on the CNS and BBB has been extensively investigated and reviewed [
239
242
].
In line with the “endotoxin hypothesis,” developing multi-organs-on-a-chip platforms
could assist in filling in the knowledge gap concerning how the microbiome may contribute
to neurodegeneration [243245].
Besides bacteria, fungi can penetrate the BBB, with Cryptococcus neoformans being the
most common one to cause meningitis. A promising NVU-on-a-chip was established by
Kim et al. [
246
], which was used to analyze the penetration of the BBB by Cryptococcus
neoformans. The pump-free model comprised a unilateral medium flow, human neural
stem cells, BMECs, and pericytes. The authors observed an elevation of inflammatory and
angiogenesis-related cytokines, such as IL-8 and thrombospondin-1, but there was no BBB
impairment. Therefore, a transcytosis-mediated entry of the pathogen into the CNS was
proposed, and the authors concluded that including additional tissues in the model to
create a multi-organ-on-chip model would be beneficial. This could help in examining the
gut–brain axis and in studying fungal penetration from the gut through the BBB, as well as
in verifying its neurotropism.
Over the last few years, the COVID-19 epidemic has led to comprehensive research on
severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2). Since neurological
symptoms were reported early on during the acute phase, but also in long-term COVID-19,
the transmigration through the BBB and the effects of a severe viral infection of the CNS
were extensively studied. Bipolar neuron-to-neuron spread in the olfactory epithelium,
transport via the vagal nerve to the brain stem, or transmigration through the blood-
cerebrospinal fluid barrier or BBB (directly or on leukocytes) were all under discussion
as entry routes for SARS-CoV-2 [
247
]. Kase et al. [
248
] observed that the pseudo-typed
lentivirus particles of major SARS-CoV-2 strains were able to infect microglia, whereas they
rarely infected other CNS cell types such as iPSC-derived neurons and astrocytes. Another
study by Andrews et al. [
249
] produced contradicting results in their stem-cell-derived
organoids, with astrocytes being the major target of SARS-CoV-2. Rhea et al. [
250
] tested the
Int. J. Mol. Sci. 2023,24, 12699 20 of 40
radiolabeled monomeric SARS-CoV-2 spike protein subunit S1 on its ability to cross the BBB
in vivo and in vitro. The authors suggested an adsorptive transcytotic manner with the par-
ticipation of angiotensin-converting enzyme 2, which is the main endogenous receptor of
the virus, in a murine brain and lung. Moreover, they noticed a more effective uptake of S1
in all brain regions when it was taken up across the BBB compared to the nasal route. This
uptake of S1 was only slightly affected by LPS-induced inflammation, whereas LPS inflam-
mation altered the S1 clearance from blood and its uptake by peripheral tissues. However,
these transwell experiments with human iBMECs showed a limited permeability for S1,
probably due to technical reasons. Buzhdygan et al. [
251
] used a hydrogel-based BBB model,
which they exposed to the SARS-CoV-2 spike protein subunits S1 and S2. Although the
spike protein induced a BBB integrity loss, it did not affect BMEC viability; rather, it en-
hanced the vascular cell/intercellular adhesion molecule-1, the pro-inflammatory response
(i.e., IL-1
β
, IL-6, C-X-C motif ligand-10, and CCL5), and MMP (especially MMP-3/12)
secretion. A study by DeOre et al. [
252
], who employed a SARS-CoV-2 spike protein in a
hydrogel-based microfluidic BBB model, revealed that the angiotensin-converting enzyme
2 expression was altered by the addition of the S1 spike protein subunit when it was paired
with fluid shear stress. Furthermore, RhoA was identified as a major regulator of the
BBB breakdown through the spike protein. Zhang et al. [
253
] found an impairment of
the BM but not TJs in their co-culture BBB transwell model with BMECs and astrocytes
after SARS-CoV-2 infection. The virus was able to replicate in the BMECs; furthermore, it
passed the barrier, probably via transcytosis, and degraded the BM (by upregulated MMP-
9). Manosso et al. [
254
] investigated the microbiota–gut–brain communication during
SARS-CoV-2 infection and suggested that the SARS-CoV-2-induced cytokine storm leads to
microglial activation, astrocyte reactivity, and neuronal degeneration, which then promote
the development of psychiatric and neurological symptoms. In a study by Ju et al. [
255
],
it was suggested that the envelope protein is responsible for the breakdown of the BBB
during SARS-CoV-2 infection. The transwell model used showed a decrease in cell viability
and an increase in inflammatory mediators (major histocompatibility complex-I, IL-1
β
,
and—particularly—IL-6) when the envelope protein was added. Furthermore, the ZO-1
mRNA levels were decreased after SARS-CoV-2 envelope protein administration.
Human immunodeficiency viruses (HIVs), different flaviviruses, and new-world al-
phaviruses have also been studied for BBB penetration and CNS infection in vitro [
201
,
256
].
Several studies showed the ability of the Zika virus and other flaviviruses in terms of
infecting and activating BMECs [
257
262
]. Moreover, the virus crossed the BBB via tran-
scytosis without significantly increasing permeability [
258
260
]. However, controversial
results from other studies have suggested that TJ breaks down in a virus strain-dependent
manner, thereby enabling for a CNS entry via a paracellular route [263]. For HIV-1, it was
demonstrated that TJ proteins are disrupted upon infection [
201
]. Exposure to a HIV-1
transactivator of transcription proteins led to ZO-1 downregulation, which was mediated
by BMEC autophagy induction in vitro [
264
,
265
]. Furthermore, the transactivator of tran-
scription was able to cross the BBB in a bidirectional manner [
266
]. Investigating the impact
of this protein on the BBB is of special importance since it affects most cell types in the CNS,
thereby contributing to neurotoxicity in HIV-1 associated neurocognitive disorders [
267
].
A recent review of Swingler et al. [
268
] highlights the applicability of the iPSC-derived
organoids of different brain regions for various neurotropic viruses. By organoid modeling,
the molecular regulation of neurotrophic viral infections, as well as cellular responses
become more accessible. However, the known limitations of organoids, such as size and
necrotic center formation, apply here as well.
5.3. Acute CNS Injuries
5.3.1. Stroke
Functional and structural BBB alterations have been observed in vivo and in vitro in
several stages of stroke. Due to the vast knowledge about the impact on the consequences
of BBB breakdown in ischemic stroke, the current BBB in vitro models focus on replicating
Int. J. Mol. Sci. 2023,24, 12699 21 of 40
neuroinflammation through oxygen-glucose deprivation (OGD). The goal of these models
is to restore the tightness of the BBB or to prevent its impairment, with the ultimate aim of
facilitating drug development [
269
273
]. Cell death due to ischemia and peripheral immune
cells causes more ROS, miRNAs, and damage-associated molecular patterns (DAMPs) to
be generated, which activate glial cells. Microglial and astrocyte activation lead to pro-
inflammatory mediator release (especially TNF-
α
and IL-1
β
), as well as the upregulation of
pro-inflammatory genes in BMECs [
12
]. MMPs, which are primarily secreted by neutrophils,
further digest TJ proteins [
274
]. Fattakhov et al. [
269
] and Kadir et al. [
270
] provided
protocols for the establishment of triple-culture models in transwell setups. Lyu et al. [
271
]
evaluated the restorative potential of stem cell therapies with their microfluidic NVU-on-a-
chip. After the functional response to OGD, the authors tracked the infiltration of candidate
stem cells through the BBB and observed that different types of stem cells exerted unique
neurorestorative effects on the structural and functional integrity of the NVU rather than
on the direct replacement of the neurons. A 3D microfluidic model by Cho et al. [
147
]
with rat brain endothelial cells was developed to screen BBB-targeting drugs. The authors
induced ischemic conditions by OGD, as well as showed the protective functions of the
antioxidant edaravone and Rho kinase-inhibitor Y-27632. Wevers et al. [
273
] modeled stroke
by introducing hypoglycemic conditions in a glucose-free medium, hypoxia by 10
µ
M of
antimycin-A (an inhibitor of complex III of the electron transport chain), and disrupted the
perfusion by moving the chips from a rocking platform to static conditions in the incubator.
Due to the relatively high-throughput BBB chip, with 40 chips in parallel, this model was
proposed as a tool for the drug screening of anti-inflammatory and free radical scavengers.
Similar drugs were utilized in the NVU organoid model of Nzou et al. [
272
] when they
were investigating their neuroprotective impacts. After inducing hypoxia by exposing the
organoids to 0.1% O
2
for 24 h, BBB breakdown, as well as pro-inflammatory cytokine and
ROS production was observed.
5.3.2. Traumatic Brain Injury
BBB impairment has been observed in all stages of traumatic brain injury (TBI), rang-
ing from mild to severe. The disruption of the BBB occurs within the first hours after TBI
due to neuroinflammation, and this may remain for years [
275
,
276
]. Despite decades of
investigations through using in vivo models, there is no efficient neuroprotective treatment
for TBI that has yet passed clinical trials. This is partly caused by the incomplete knowledge
about molecular mechanisms in the complex pathophysiology of TBI. Therefore, in vitro
models could be helpful in investigating the underlying structural and functional alter-
ations at the cellular level. Different methods have been employed to induce neurotrauma
in vitro since TBI itself can also result from different impacts, such as blunt-force, blast,
or compression [
277
]. These injury models are induced statically or dynamically by me-
chanical or chemical forces. Static mechanical injuries are impact-based methods such
as weight drop. Dynamic methods include microfluidic compartmentalization or scratch
assays, as well as stretch-induced traumas that are caused by culture monolayer or axonal
stretching. Chemical injuries are used for the microenvironment investigations of post-TBI
tissues, and these include altering culture conditions with nutrient or oxygen deprivation,
or treatments with chemicals [
278
]. Schlotterose et al. [
279
] developed a 3D-printed device
for standard and non-standard cell culture applications to induce hydrostatic pressure
on cells, resulting in their perturbation. As such, they were able to provoke TBI-typical
hallmarks, such as cell death, decreased neuronal functionality, neural axon swelling, and a
decrease in BBB tightness. Uni- and biaxial cell stretch models were set up by Rosas-
Hernandez et al. [
280
,
281
] to assess cell viability and impact on the BBB after mild TBI.
The model was based on rat BMECs on a silicone membrane, which was stretched utilizing
a custom-made stretching device. They observed a deformation-dependent increase in
cell death and apoptosis after high magnitudes of stretch, whereas the metabolic activity
in the stretched BMECs was already decreased after a lower magnitude of deformation.
Interestingly, low-magnitude stretching enhanced TJ protein expression, indicating its po-
Int. J. Mol. Sci. 2023,24, 12699 22 of 40
tentially protective role in BBB integrity. However, the authors also argued that experiments
should be repeated with human BMECs since their results were shown to be contrary to the
results of other studies that used mouse BMECs instead, thus indicating species-to-species
differences. Salvador et al. [
282
] provided a protocol through which to perform a stretch-
and OGD-induced in vitro model that could be used to mimic the impact of TBI on mouse
BMECs. This approach can also be performed with human BMECs so as to serve a more
physiologically relevant model. A systemic review by Wu et al. [
283
] summarized the
current, at the time, in vitro models for TBI up until 2021. Only a very small proportion
included investigations of the BBB alterations during trauma. Therefore, future studies are
needed to reveal the impact of BBB dysregulation in the different forms of TBI. Furthermore,
other cell types than solely BMECs, astrocytes, or neurons should be included in the setups.
For example, pericytes were shown to rapidly detach from BMECs after TBI in a mouse
model, but they were significantly enhanced again after 5 days, which indicates a biphasic
pericyte regulation in acute TBI [284].
5.3.3. Epilepsy
For decades, it has been discussed if BBB leakage is the cause or consequence for
seizure development and aggravation [
203
,
285
]. A major problem in the treatment of
epilepsy is the development of drug resistance in a third of patients. This treatment
failure could be caused by an excessive upregulation of efflux transporters, such as P-gp or
the molecular target that are changes caused by anti-epileptic drugs themselves [
47
,
286
].
To study the interactions of epileptic tissue and BBB, several in vitro models have been
developed. However, although being labeled as in vitro models, many of these models have
been of organotypic brain slices or whole brains that were isolated from rodents [
287
289
].
Seizure-like events can be induced by low magnesium concentrations, multiple doses of
kainic acid, or 4-aminopyridine (a Ca
2+
channel blocker) [
290
]. To achieve personalized
drug development for patients, the organotypic cultures of tissue obtained from epileptic
patients can be isolated and used in co-culture contact or in non-contact transwell models
along with BMECs, as well as in brain tissues that exhibit seizure-like activities [
291
].
This allows one to display TJ and adhesion protein expressions in the monolayer, as well
as to investigate the bidirectional interactions of BMECs with epileptic tissue, especially
after the application of different anti-epileptic drugs [
286
]. So far, there are no in vitro
models of epilepsy that incorporate more BMECs and astrocytes/isolated brain tissue [
200
].
Furthermore, working with isolated human brain tissue is limited due to high complexity
and low availability [
292
]. Yamanaka et al. [
293
] reviewed the neuroinflammatory role
of pericytes and their important impact in epilepsy pathogenesis. Pericytes undergo
redistribution and remodeling during epileptic events, thereby causing BBB alterations due
to responses to pro-inflammatory cytokines, as well as pericyte-glial scarring at permeable
microvessels. Therefore, including pericytes in epilepsy in vitro co-culture models would
possibly provide more physiologically reliable results for human drug discovery [
122
,
294
].
5.4. CNS Tumors
Brain cancers display significant heterogeneity, arising either from the CNS itself or
due to spreading from the periphery as metastases through a permeable BBB. The BBB ex-
periences several alterations during tumor cell invasion. This is also often referred to as the
blood–brain tumor barrier, as microvessels have highly heterogeneous permeability follow-
ing the active efflux of molecules, enhanced angiogenesis, and inflammation when they are
combined with less blood flow due to tumor growth. The highest permeability is frequently
found at lesion cores [
200
]. Novel-formed microvessels within the tumor tissue exhibit
neither sufficient intercellular junctions nor transport systems [
47
]. With tumor progression,
astrocytes lose their connection to the vasculature and interact with tumor cells, thereby
regulating proliferation, immune cell invasion, and drug responses [
295
]. Furthermore,
edemas due to AQP-4 upregulation is regularly observed in brain tumors [
12
]. Otherwise,
BBB permeability in brain tumors is a two-edged sword. Although the barrier is leaky,
Int. J. Mol. Sci. 2023,24, 12699 23 of 40
chemotherapeutics either cannot enter, because the breakdown of the BBB is unequally
distributed, or they are removed again from the CNS by active efflux mechanisms [
296
].
Researchers have already attempted to exploit neuroinflammation in order to open the
BBB. Blethen et al. [
296
] provide an overview of methods, such as low-intensity-focused
ultrasound, which can be employed to modulate the blood–brain tumor barrier in order to
enhance drug uptake. These methods induce a secondary inflammation process that leads
to the transient intercellular opening of BMECs and is mediated by the elevated release of
DAMPs, TNF-
α
, IL-8, and heat shock proteins into the brain parenchyma. Although this
approach was already proven to be successful in some preclinical studies, there is more
research needed for the fine-tuning of low-intensity-focused ultrasound. Tumor modeling
benefits greatly from 3D in vitro modeling as the tumor microenvironment can be mimicked
for chemotherapeutic screening and metastasis studies. Experimental setups include multi-
ple cell types in 3D scaffolds, thereby adding tumor spheroids in the brain parenchyma
or vessel compartment [
297
300
]. Seo et al. [
301
] established a glioblastoma multiforme
spheroid, which they placed in a vascularized 3D hydrogel to study chemosensitivity and
the drug delivery associated with the BBB in this tumor. They added TNF-
α
and saw
enhanced BBB permeability and monocytic THP-1 cell adhesion to BMECs due to higher
adhesion molecule expression. The authors also highlighted the importance of other NVU
cell types, such as astrocytes for BBB integrity, but also their prohibitive role in cancer metas-
tasis. A recent study by Zhang et al. [
302
] suggested the enhanced adhesion of circulating
breast cancer cells with metastasis potential to the brain endothelium. The microfluidic
model included varying shear stress, selecting those cells with the highest adhesion to
BMECs. As such, the authors observed not only elevated brain metastasis gene expressions
in circulating breast cancer cells, but also increased transmigrations through the BBB in
a transwell model. Moreover, their adhesion and proliferation within polyacrylamide
gels (0.6 kPa) (which were coated with collagen I) was better than the one of wild-type
cancer cells, which the research group associated with a better survival in the soft brain
microenvironment. Finally, immunocompetent glioblastoma organoids could bring about
a breakthrough in personalized medicine when co-cultured with patient blood/tumor-
derived immune cells for the development of immunotherapies [
303
]. One patient-specific
glioblastoma model was established by Cui et al. [
304
], who aimed to optimize patient-
specific responses to programmed cell death protein-1 checkpoint immunotherapy by
dissecting the immunosuppressive tumor microenvironment heterogeneity.
5.5. Neurodegeneration
5.5.1. Alzheimer’s Disease
Independently of whether it is the cause or consequence, the involvement of BBB
disruption in AD has been proven to a large extent [
305
]. Evidence for the BBB breakdown
in AD patients also comes from plasma proteins like fibrin(ogen), thrombin, albumin, and
the immunoglobulins that are found co-localized with A
β
plaques [
16
,
306
308
]. In AD, glial
cells are primarily engaged in neuroinflammation. Studies have shown the activated astro-
cytes and microglia around plaques, which release pro-inflammatory cytokines and trigger
further inflammatory processes. Astrocyte alterations besides astrogliosis in AD comprise
an impairment in glucose metabolism, a decreased expression of glutamate transporters,
and an imbalance in potassium [
309
]. A recent small nuclear RNA study by Xu et al. [
310
] of
human postmortem brains revealed the transcriptomic changes in astrocytes and microglia
in pathogenic conditions. Although astrocytic changes had common features in AD and
PD, the regional differences were provocative toward one of the pathologies. In contrast,
the microglia showed unique disorder gene transcriptomic patterns. In vitro BBB models
of AD incorporate inflammatory regulation by adding pro-inflammatory cytokines and A
β
.
Spampinato et al. [
305
] investigated the peripheral blood mononuclear cell transmigration
across the barrier under A
β
exposure and pro-inflammatory conditions, as well as provided
a protocol for a transwell model setup. Schreiner et al. [
311
] summarized the current AD
BBB in vitro models in great detail.
Int. J. Mol. Sci. 2023,24, 12699 24 of 40
5.5.2. Amyotrophic Lateral Sclerosis
Blood-derived components like the thrombin, immunoglobulin G, or hemoglobin that
have been found in ALS patient postmortem brains indicate BBB breakdown. Furthermore,
a decrease in TJ protein expression, astrogliosis with detached astrocyte endfeet, and severe
pericyte degradation have also been observed [
16
,
312
]. In ALS progression, adaptive
immunity may contribute to the disease severity, and recent studies have highlighted the
CNS infiltration of T cells, as well as an overall shift toward pro-inflammatory T cell subsets
and reduced CD4+ T cells [
313
]. For example, CD4+ Tregs are reduced during disease
progression. Their suppressive capability is also diminished in vitro by isolating patient T
cells. Interestingly, removing Tregs from their environment could restore this characteristic
feature, thererby displaying a possible therapeutic target for the autologous transplantation
of T cells [
314
]. Most studies regarding BBB breakdown in ALS have been performed in
SOD1 mutant rodent models or in postmortem human tissues [
315
,
316
]. In vitro, not many
models have been established until recently. The majority of models have been of the
cell cultures of motor neuron cell lines with the abovementioned mutations, or of patient-
derived motor neurons. Therefore, the focus has been to understand cellular mechanisms
more specifically. In recent years, there has been a significant focus on using iPSCs to better
understand the causes of ALS and to develop effective treatment approaches [
317
]. A recent
review by Arjmand et al. [
318
] provides an input about how an ALS organ-on-a-chip can
be approached. Embryonic or neural stem cells or iPSCs can be derived from patient
fibroblasts. The microenvironment should involve an excess of glutamic acid to mimic the
glutamate-induced excitotoxicity.
5.5.3. Parkinson’s Disease
α
-Synuclein acts in a pro-inflammatory manner via different mechanisms in PD [
81
].
On the one hand, it binds to microglia and astrocytes via toll-like receptor-2 and -4, thus
leading to the release of inflammatory mediators [
319
,
320
]. On the other hand, monomeric
α
-synuclein was shown to induce the release of pro-inflammatory cytokines/chemokines
by pericytes in a transwell model, thus leading to reduced BMEC integrity [
321
]. En-
hanced BBB impairment was also seen in PD patients through using quantitative MRI
imaging [
322
]. However, although pre-formed
α
-synuclein fibrils lead to a downregu-
lation of TJ proteins, BBB breakdown seems to be dependent on multiple inflammatory
events since the functionality of the BBB by the addition of
α
-synuclein alone was not
altered significantly in vitro [
323
]. Therefore, including multiple cell types to model PD
is crucial for reflecting the pathology, and for performing reliable experiments. Pediadi-
takis et al. [
324
] established a “human Substantia Nigra Brain-Chip” consisting of a PDMS
chip with two perfusable microchannels and an ECM component mixture-coated PDMS
membrane in between. Thus, they introduced iPSC-derived dopaminergic neurons, human
primary astrocytes, pericytes, and microglia into one of the two channels, and iPSC-derived
BMECS on the surface of the opposite channel. Continuous exposure of
α
-synuclein fibrils
in the “brain” channel showed an accumulation of pSer129-
α
Syn fibrils, mitochondrial
impairment, and an enhanced BBB permeability after a few days.
5.5.4. Huntington’s Disease
A study by Vignone et al. [
109
] compared the iBMECs of HD patients with those of
the unaffected controls. They aimed to investigate whether changes in the transcriptomic
profiles indicated if the BMECs themselves were functionally compromised when pro-
moting BBB dysfunction. The results indicated alterations in the BBB properties, and in
functions such as receptor-mediated transcytosis. A gene expression analysis of iBMECs
by Linville et al. [
325
] showed similar results regarding TJ proteins, while they could not
observe the impaired paracellular permeability in their transwell setup. Furthermore,
the authors saw differences in their immune cell adhesion and immune activation tran-
scripts between the juvenile HD iBMECs and adult postmortem HD BMECs. Therefore,
they suggested BBB gene expression changes due to the elevated CAG repeat expansion
Int. J. Mol. Sci. 2023,24, 12699 25 of 40
during HD progression. Besides these studies using iBMEC monocultures, not many in-
vestigations have been conducted on the differentiation of other HD-relevant cell types.
However, this could be due to the difficulty of replicating differentiation protocols (for in-
stance, striatal medium spiny neurons [
326
]). There are still many tasks that need to be
completed in order to accurately model the BBB for HD. One key area that needs develop-
ment is the creation of an in vitro co-culture model that includes other CNS resident cells
to better determine the BBB phenotype.
6. Discussion
The barrier function of the BBB is strictly regulated by various cell types in the CNS.
While in the past BMECs have primarily been employed for in vitro modeling, investiga-
tions on the surrounding astrocytes, pericytes, microglia, neurons, and oligodendrocytes
have shed new light on the importance of these CNS-resident cells. In neuroinflammation,
both central and peripheral cells produce pro-inflammatory mediators that lead to BBB
impairment. Inflammation-inducing factors vary across the spectrum of neurological dis-
eases, such as the production of autoantibodies, the release of ROS and DAMPS due to
ischemia and trauma, and the influence of metastases on the microvascular efflux pumps
and permeability. While a huge variety of different models exist for some neuropathologies,
there is a devastating lack of in vitro models for others. Some disease models could bene-
fit from the already-established setups of other diseases. While monoculture transwells
may not be suitable for more complex physiological research questions or personalized
approaches, they still offer advantages in conducting basic high-throughput toxicity or
permeability investigations. Models that incorporate multiple cell types better mimic
human (patho)physiology, even more so when applied to a 3D model setup with an ECM
substitute and shear stress. The tunability of the incorporated cells can be established by
using patient-derived iPSCs to enable more personalized research regarding drug tolerance,
toxicity, and efficacy. It is worth noting that as the model becomes more complex, there are
more potential sources of error that can accumulate. To validate the model, the cell-type
identity and viability need to be confirmed. Permeability can be measured by TEER, or
via size-defined particles that can pass a leaky barrier. Appropriate barrier formation is
evaluated via TJ protein detection. For the future, a robust and reliable platform mimicking
different neurological diseases that are as close to in vivo conditions as possible is desirable
in order to reduce lab-to-lab aberrancies. Although in vitro models will never fully reflect
human physiology, they are great tools that allow for increasing reduction and replacement
of animal models.
7. Conclusions
There are several key considerations for BBB in vitro modeling that have been ad-
dressed in this review. First, considerations need to be made regarding the “why” of the
experiment. The model setup can completely differ depending on whether the transport
of substances across a BMEC monolayer is examined, or how the tumor microenviron-
ment in glioblastoma organoids changes due to different immunotherapy approaches.
Each pathology brings different characteristics and biological requirements that need to
be adapted accordingly in the BBB model. For example, the majority of patients with
NMOSD exhibit AQP-4 antibody seropositivity. To investigate the pathomechanism of
these antibodies against astrocytes, it is essential to evaluate the AQP-4 expression of these
cells before conducting experiments. Investigations of AD or PD need the presence of A
β
or
α
-synuclein fibrils, respectively, on the “brain” side. Acute CNS injuries need to be induced
before the investigations, such as OGD in ischemia models or those in the differently
provoked injuries in TBI. Additionally, when modeling neuroinflammation, it is important
to choose a method for inducing inflammation. Most models aim for stimulation with
LPS or TNF-
α
, but the addition of other pro-inflammatory substances can be considered as
well. Moreover, the cell and ECM types, as well as the general setup (simple BBB vs. full
NVU), need to be considered. For experiments on the function of the different transporters
Int. J. Mol. Sci. 2023,24, 12699 26 of 40
or receptors of BMECs, a BMEC monoculture model that shows the respective protein
expression can be sufficient. However, investigations of the different cell types’ interactions
might require more complex setups. For transwell models, the multi-well format, pore size,
and membrane material, as well as a BM substitute need to be defined. For hydrogel-based
models, the hydrogel composition and cross-linking substance need to be evaluated; in
addition, some additives might be required for enhanced cell viability. Organoids need a
cocktail of growth factors and an ECM for proper differentiation. Furthermore, the time
frame of each experimental approach may vary. Utilizing iPSCs for the model can be
time-consuming since the differentiation protocols require several weeks to complete. Then,
there is a short time window where the differentiated cells can be used for the model
before de-differentiation. Primary cells can be commercially purchased or isolated directly
(if ethically applicable), but they can only be used for a few passages, which limits the
number of attempts and cells for the model establishment. In contrast, immortalized cell
lines can be used over many passages in higher quantities. Moreover, in particular for
drug safety studies in pharmaceutical production, a high-throughput platform is essential.
Often, the quantity of the model compromises its quality or, in this case, its complexity.
Hence, high-throughput models most likely will not require time-intensive fabrication.
For some models, additional devices are needed. For the introduction of shear stress, a per-
fusion system is required. This can involve different pumping systems or an incubator-safe
rocking platform. BBB/NVU-on-a-chip models need specially fabricated chips made of
PDMS or similar materials. It should be noted that there are various printing techniques
used for 3D bioprinted models that require different devices based on the printing material
and complexity of the model. To measure the TEER, specialized TEER meters need to be
purchased. Immunofluorescence staining requires fluorescence microscopes. If experiments
are conducted in 3D setups, a laser confocal microscope can be beneficial for creating a
3D representation of the staining. Finally, the culture time is dependent on the cells used
and the experimental procedure performed. Experiments are usually started when BMECs
have been cultured long enough to develop a tight monolayer. The formation of TJs and
low permeability need to be assessed by TJ marker immunofluorescence staining, TEER
measurements, or permeability assays. The enhancement of TJs can be reached by the
addition of glucocorticoids, but this should be avoided in inflammation-related studies.
Additional incubation time might be necessary, if inflammation needs to be provoked or
if the effects of different drugs, cytokines, ROS, antibodies, or immune cells on the BBB are
being studied.
Ultimately, the selection of a suitable model setup and its complexity depend on the
specific research question at hand. As the British statistician George E. P. Box once said
(which is also applicable for biological models): “All models are wrong, but some are
useful” [327].
Funding: Supported by the intramural funding program of the Medical University Innsbruck Ph.D.
Research Training Groups, Project 2022-1-2 “CONNECT”.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
Abbreviations
The following abbreviations are used in this manuscript:
AβAmyloid-beta
AD Alzheimer’s disease
ALS Amyotrophic lateral sclerosis
AQP-4 Aquaporin-4
BBB Blood–brain barrier
Int. J. Mol. Sci. 2023,24, 12699 27 of 40
BM Basement membrane
BMEC Brain microvascular endothelial cell
CCL C-C motif ligand
CNS Central nervous system
DAMP Danger-associated molecular pattern
ECM Extracellular matrix
GABA Gamma-aminobutyric acid
HIV Human immunodeficiency virus
IL Interleukin
iBMEC Induced brain microvascular endothelial cell-like cell
iPSC Induced pluripotent stem cell
LPS Lipopolysaccharide
MMP Matrix metalloproteinase
MOG Myelin oligodendrocyte glycoprotein
MOGAD MOG-associated antibody disease
MS Multiple sclerosis
NMDARE N-methyl-D-aspartate receptor encephalitis
NMOSD Neuromyelitis optica spectrum disorders
NVU Neurovascular unit
OGD Oxygen-glucose deprivation
OPC Oligodendrocyte progenitor cell
P-gp P-glycoprotein
PD Parkinson’s disease
PDMS Polydimethylsiloxane
ROS Reactive oxygen species
SARS-CoV-2 Severe acute respiratory syndrome coronavirus type 2
TBI Traumatic brain injury
TEER Trans-endothelial electrical resistance
TGF-βTransforming growth factor-beta
Th T helper
TJ Tight junction
TNF-αTumor necrosis factor-alpha
VEGF Vascular endothelial growth factor
ZO-1 Zonula occludens-1
References
1. Pardridge, W.M. Blood-brain barrier delivery. Drug Discov. Today 2007,12, 54–61. [CrossRef] [PubMed]
2.
Banks, W.A.; Greig, N.H. Small molecules as central nervous system therapeutics: Old challenges, new directions, and a
philosophic divide. Future Med. Chem. 2019,11, 489–493. [CrossRef] [PubMed]
3.
Nag, S.; David, J.B. Blood Brain Barrier, Exchange of metabolites and gases. In Pathology and Genetics: Cerebrovascular Diseases;
Kalimo, H., Ed.; ISN Neuropath Press: Basel, Switzerland, 2005; pp. 22–29.
4.
Gifre-Renom, L.; Daems, M.; Luttun, A.; Jones, E.A.V. Organ-Specific Endothelial Cell Differentiation and Impact of Microenvi-
ronmental Cues on Endothelial Heterogeneity. Int. J. Mol. Sci. 2022,23, 1477. [CrossRef] [PubMed]
5.
Yang, Q.; Wijerathne, H.; Langston, J.C.; Kiani, M.F.; Kilpatrick, L.E. Emerging Approaches to Understanding Microvascular
Endothelial Heterogeneity: A Roadmap for Developing Anti-Inflammatory Therapeutics. Int. J. Mol. Sci.
2021
,22, 7770.
[CrossRef]
6.
Abbott, N.J.; Patabendige, A.A.K.; Dolman, D.E.M.; Yusof, S.R.; Begley, D.J. Structure and function of the blood-brain barrier.
Neurobiol. Dis. 2010,37, 13–25. [CrossRef]
7. Stamatovic, S.M.; Johnson, A.M.; Keep, R.F.; Andjelkovic, A.V. Junctional proteins of the blood-brain barrier: New insights into
function and dysfunction. Tissue Barriers 2016,4, e1154641. [CrossRef]
8.
Armulik, A.; Genové, G.; Betsholtz, C. Pericytes: Developmental, physiological, and pathological perspectives, problems, and
promises. Dev. Cell 2011,21, 193–215. [CrossRef]
9. Simon, A.M.; Goodenough, D.A. Diverse functions of vertebrate gap junctions. Trends Cell Biol. 1998,8, 477–483. [CrossRef]
10.
Furuse, M.; Sasaki, H.; Tsukita, S. Manner of interaction of heterogeneous claudin species within and between tight junction
strands. J. Cell Biol. 1999,147, 891–903. [CrossRef]
11.
Hashimoto, Y.; Greene, C.; Munnich, A.; Campbell, M. The CLDN5 gene at the blood-brain barrier in health and disease. Fluids
Barriers CNS 2023,20, 22. [CrossRef]
12.
Kadry, H.; Noorani, B.; Cucullo, L. A blood-brain barrier overview on structure, function, impairment, and biomarkers of
integrity. Fluids Barriers CNS 2020,17, 69. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 28 of 40
13.
Pardridge, W.M. Drug transport across the blood-brain barrier. J. Cereb. Blood Flow Metab.
2012
,32, 1959–1972. [CrossRef]
[PubMed]
14. Banks, W.A. Characteristics of compounds that cross the blood-brain barrier. BMC Neurol. 2009,9(Suppl. 1), S3. [CrossRef]
15.
Taylor, C.J.; Nicola, P.A.; Wang, S.; Barrand, M.A.; Hladky, S.B. Transporters involved in regulation of intracellular pH in primary
cultured rat brain endothelial cells. J. Physiol. 2006,576, 769–785. [CrossRef]
16.
Sweeney, M.D.; Zhao, Z.; Montagne, A.; Nelson, A.R.; Zlokovic, B.V. Blood-Brain Barrier: From Physiology to Disease and Back.
Physiol. Rev. 2019,99, 21–78. [CrossRef]
17.
Shibata, M.; Yamada, S.; Kumar, S.R.; Calero, M.; Bading, J.; Frangione, B.; Holtzman, D.M.; Miller, C.A.; Strickland, D.K.; Ghiso,
J.; et al. Clearance of Alzheimer’s amyloid-ss(1-40) peptide from brain by LDL receptor-related protein-1 at the blood-brain
barrier. J. Clin. Investig. 2000,106, 1489–1499. [CrossRef]
18.
Zhao, Z.; Nelson, A.R.; Betsholtz, C.; Zlokovic, B.V. Establishment and Dysfunction of the Blood-Brain Barrier. Cell
2015
,
163, 1064–1078. [CrossRef]
19.
Baeten, K.M.; Akassoglou, K. Extracellular matrix and matrix receptors in blood-brain barrier formation and stroke. Dev.
Neurobiol. 2011,71, 1018–1039. [CrossRef]
20.
Tietz, S.; Engelhardt, B. Brain barriers: Crosstalk between complex tight junctions and adherens junctions. J. Cell Biol.
2015
,
209, 493–506. [CrossRef]
21.
Jin, J.; Fang, F.; Gao, W.; Chen, H.; Wen, J.; Wen, X.; Chen, J. The Structure and Function of the Glycocalyx and Its Connection
With Blood-Brain Barrier. Front. Cell. Neurosci. 2021,15, 739699. [CrossRef]
22.
Salmon, A.H.J.; Satchell, S.C. Endothelial glycocalyx dysfunction in disease: Albuminuria and increased microvascular perme-
ability. J. Pathol. 2012,226, 562–574. [CrossRef]
23.
Lyu, N.; Du, Z.; Qiu, H.; Gao, P.; Yao, Q.; Xiong, K.; Tu, Q.; Li, X.; Chen, B.; Wang, M.; et al. Mimicking the Nitric Oxide-Releasing
and Glycocalyx Functions of Endothelium on Vascular Stent Surfaces. Adv. Sci. 2020,7, 2002330. [CrossRef] [PubMed]
24.
Zhao, F.; Zhong, L.; Luo, Y. Endothelial glycocalyx as an important factor in composition of blood-brain barrier. CNS Neurosci.
Ther. 2021,27, 26–35. [CrossRef] [PubMed]
25.
Bush, M.A.; Anstey, N.M.; Yeo, T.W.; Florence, S.M.; Granger, D.L.; Mwaikambo, E.D.; Weinberg, J.B. Vascular Dysfunction in
Malaria: Understanding the Role of the Endothelial Glycocalyx. Front. Cell Dev. Biol. 2021,9, 751251. [CrossRef] [PubMed]
26.
Logsdon, A.F.; Rhea, E.M.; Reed, M.; Banks, W.A.; Erickson, M.A. The neurovascular extracellular matrix in health and disease.
Exp. Biol. Med. 2021,246, 835–844. [CrossRef] [PubMed]
27. Allt, G.; Lawrenson, J.G. Pericytes: Cell biology and pathology. Cells Tissues Organs 2001,169, 47855. [CrossRef] [PubMed]
28. Sims, D.E. Recent advances in pericyte biology–implications for health and disease. Can. J. Cardiol. 1991,7, 431–443.
29.
Longden , T.A.; Zhao, G.; Hariharan, A.; Lederer, W.J. Pericytes and the Control of Blood Flow in Brain and Heart. Annu. Rev.
Physiol. 2023,85, 137–164. [CrossRef]
30.
Winkler, E.A.; Bell, R.D.; Zlokovic, B.V. Central nervous system pericytes in health and disease. Nat. Neurosci.
2011
,14, 1398–1405.
[CrossRef]
31.
Lindahl, P.; Johansson, B.R.; Levéen, P.; Betsholtz, C. Pericyte loss and microaneurysm formation in PDGF-B-deficient mice.
Science 1997,277, 242–245. [CrossRef]
32.
Dehouck, M.P.; Vigne, P.; Torpier, G.; Breittmayer, J.P.; Cecchelli, R.; Frelin, C. Endothelin-1 as a mediator of endothelial
cell-pericyte interactions in bovine brain capillaries. J. Cereb. Blood Flow Metab. 1997,17, 464–469. [CrossRef] [PubMed]
33.
van Zwieten, E.J.; Ravid, R.; Swaab, D.F.; van de Woude, T. Immunocytochemically stained vasopressin binding sites on blood
vessels in the rat brain. Brain Res. 1988,474, 369–373. [CrossRef] [PubMed]
34.
Healy, D.P.; Wilk, S. Localization of immunoreactive glutamyl aminopeptidase in rat brain. II. Distribution and correlation with
angiotensin II. Brain Res. 1993,606, 295–303. [CrossRef] [PubMed]
35. Freeman, M.R. Specification and Morphogenesis of Astrocytes. Science 2010,330, 774–778. [CrossRef] [PubMed]
36.
Seifert, G.; Schilling, K.; Steinhäuser, C. Astrocyte dysfunction in neurological disorders: A molecular perspective. Nat. Rev.
Neurosci. 2006,7, 194–206. [CrossRef]
37.
Abbott, N.J.; Rönnbäck, L.; Hansson, E. Astrocyte-endothelial interactions at the blood-brain barrier. Nat. Rev. Neurosci.
2006
,
7, 41–53. [CrossRef]
38.
Nielsen, S.; Nagelhus, E.A.; Amiry-Moghaddam, M.; Bourque, C.; Agre, P.; Ottersen, O.P. Specialized membrane domains
for water transport in glial cells: High-resolution immunogold cytochemistry of aquaporin-4 in rat brain. J. Neurosci.
1997
,
17, 171–180. [CrossRef]
39.
Hibino, H.; Inanobe, A.; Furutani, K.; Murakami, S.; Findlay, I.; Kurachi, Y. Inwardly rectifying potassium channels: Their
structure, function, and physiological roles. Physiol. Rev. 2010,90, 291–366. [CrossRef]
40.
Ohno, Y.; Kinboshi, M.; Shimizu, S. Inwardly Rectifying Potassium Channel Kir4.1 as a Novel Modulator of BDNF Expression in
Astrocytes. Int. J. Mol. Sci. 2018,19, 3313. [CrossRef]
41.
Almad, A.; Maragakis, N.J. A stocked toolbox for understanding the role of astrocytes in disease. Nat. Rev. Neurol.
2018
,
14, 351–362. [CrossRef]
42.
Lee, H.G.; Wheeler, M.A.; Quintana, F.J. Function and therapeutic value of astrocytes in neurological diseases. Nat. Rev. Drug
Discov. 2022,21, 339–358. [CrossRef] [PubMed]
43. Sofroniew, M.V. Astrocyte barriers to neurotoxic inflammation. Nat. Rev. Neurosci. 2015,16, 249–263. [CrossRef] [PubMed]
Int. J. Mol. Sci. 2023,24, 12699 29 of 40
44.
Sofroniew, M.V. Astrocyte Reactivity: Subtypes, States, and Functions in CNS Innate Immunity. Trends Immunol.
2020
,41, 758–770.
[CrossRef] [PubMed]
45.
Kaplan, L.; Chow, B.W.; Gu, C. Neuronal regulation of the blood–brain barrier and neurovascular coupling. Nat. Rev. Neurosci.
2020,21, 416–432. [CrossRef] [PubMed]
46.
Schreiner, T.G.; Romanescu, C.; Popescu, B.O. The Blood-Brain Barrier-A Key Player in Multiple Sclerosis Disease Mechanisms.
Biomolecules 2022,12, 538. [CrossRef]
47. Sommonte, F.; Arduino, I.; Racaniello, G.F.; Lopalco, A.; Lopedota, A.A.; Denora, N. The Complexity of the Blood-Brain Barrier
and the Concept of Age-Related Brain Targeting: Challenges and Potential of Novel Solid Lipid-Based Formulations. J. Pharm.
Sci. 2022,111, 577–592. [CrossRef]
48.
Pulido, R.S.; Munji, R.N.; Chan, T.C.; Quirk, C.R.; Weiner, G.A.; Weger, B.D.; Rossi, M.J.; Elmsaouri, S.; Malfavon, M.; Deng, A.;
et al. Neuronal Activity Regulates Blood-Brain Barrier Efflux Transport through Endothelial Circadian Genes. Neuron
2020
,
108, 937–952.e7. [CrossRef]
49.
Kwon, H.S.; Koh, S.H. Neuroinflammation in neurodegenerative disorders: The roles of microglia and astrocytes. Transl.
Neurodegener. 2020,9, 42. [CrossRef]
50.
Maguire, E.; Connor-Robson, N.; Shaw, B.; O’Donoghue, R.; Stöberl, N.; Hall-Roberts, H. Assaying Microglia Functions In Vitro.
Cells 2022,11, 3414. [CrossRef]
51.
Subhramanyam, C.S.; Wang, C.; Hu, Q.; Dheen, S.T. Microglia-mediated neuroinflammation in neurodegenerative diseases.
Semin. Cell Dev. Biol. 2019,94, 112–120. [CrossRef]
52.
Pfeiffer, F. Reciprocal Interactions between Oligodendrocyte Precursor Cells and the Neurovascular Unit in Health and Disease.
Cells 2022,11, 1954. [CrossRef] [PubMed]
53.
de La Fuente, A.G.; Lange, S.; Silva, M.E.; Gonzalez, G.A.; Tempfer, H.; van Wijngaarden, P.; Zhao, C.; Di Canio, L.; Trost, A.;
Bieler, L.; et al. Pericytes Stimulate Oligodendrocyte Progenitor Cell Differentiation during CNS Remyelination. Cell Rep.
2017
,
20, 1755–1764. [CrossRef] [PubMed]
54.
Cherry, J.D.; Olschowka, J.A.; O’Banion, M.K. Neuroinflammation and M2 microglia: The good, the bad, and the inflamed. J.
Neuroinflamm. 2014,11, 98. [CrossRef]
55.
Kalyan, M.; Tousif, A.H.; Sonali, S.; Vichitra, C.; Sunanda, T.; Praveenraj, S.S.; Ray, B.; Gorantla, V.R.; Rungratanawanich, W.;
Mahalakshmi, A.M.; et al. Role of Endogenous Lipopolysaccharides in Neurological Disorders. Cells
2022
,11, 4038. [CrossRef]
[PubMed]
56.
Tang, Y.; Le, W. Differential Roles of M1 and M2 Microglia in Neurodegenerative Diseases. Mol. Neurobiol.
2016
,53, 1181–1194.
[CrossRef]
57.
Liddelow, S.A.; Guttenplan, K.A.; Clarke, L.E.; Bennett, F.C.; Bohlen, C.J.; Schirmer, L.; Bennett, M.L.; Münch, A.E.; Chung, W.S.;
Peterson, T.C.; et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 2017,541, 481–487. [CrossRef]
58.
Yu, G.; Zhang, Y.; Ning, B. Reactive Astrocytes in Central Nervous System Injury: Subgroup and Potential Therapy. Front. Cell.
Neurosci. 2021,15, 792764. [CrossRef]
59.
Escartin, C.; Galea, E.; Lakatos, A.; O’Callaghan, J.P.; Petzold, G.C.; Serrano-Pozo, A.; Steinhäuser, C.; Volterra, A.; Carmignoto,
G.; Agarwal, A.; et al. Reactive astrocyte nomenclature, definitions, and future directions. Nat. Neurosci.
2021
,24, 312–325.
[CrossRef]
60.
Hasel, P.; Rose, I.V.L.; Sadick, J.S.; Kim, R.D.; Liddelow, S.A. Neuroinflammatory astrocyte subtypes in the mouse brain. Nat.
Neurosci. 2021,24, 1475–1487. [CrossRef]
61.
Berriat, F.; Lobsiger, C.S.; Boillée, S. The contribution of the peripheral immune system to neurodegeneration. Nat. Neurosci.
2023
,
26, 942–954. [CrossRef]
62. Croese, T.; Castellani, G.; Schwartz, M. Immune cell compartmentalization for brain surveillance and protection. Nat. Immunol.
2021,22, 1083–1092. [CrossRef]
63.
Marchetti, L.; Engelhardt, B. Immune cell trafficking across the blood-brain barrier in the absence and presence of neuroinflam-
mation. Vasc. Biol. 2020,2, H1–H18. [CrossRef] [PubMed]
64.
Prinz, M.; Priller, J. The role of peripheral immune cells in the CNS in steady state and disease. Nat. Neurosci.
2017
,20, 136–144.
[CrossRef] [PubMed]
65.
Becher, B.; Spath, S.; Goverman, J. Cytokine networks in neuroinflammation. Nat. Rev. Immunol.
2017
,17, 49–59. [CrossRef]
[PubMed]
66. Rustenhoven, J.; Kipnis, J. Brain borders at the central stage of neuroimmunology. Nature 2022,612, 417–429. [CrossRef]
67.
Castellani, G.; Croese, T.; Peralta Ramos, J.M.; Schwartz, M. Transforming the understanding of brain immunity. Science
2023
,
380, eabo7649. [CrossRef]
68.
Zipp, F.; Bittner, S.; Schafer, D.P. Cytokines as emerging regulators of central nervous system synapses. Immunity
2023
,56, 914–925.
[CrossRef] [PubMed]
69.
Vafadari, B.; Salamian, A.; Kaczmarek, L. MMP-9 in translation: From molecule to brain physiology, pathology, and therapy. J.
Neurochem. 2016,139 (Suppl. 2), 91–114. [CrossRef] [PubMed]
70.
Rempe, R.G.; Hartz, A.M.S.; Bauer, B. Matrix metalloproteinases in the brain and blood-brain barrier: Versatile breakers and
makers. J. Cereb. Blood Flow Metab. 2016,36, 1481–1507. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 30 of 40
71.
Candelario-Jalil, E.; Yang, Y.; Rosenberg, G.A. Diverse roles of matrix metalloproteinases and tissue inhibitors of metallopro-
teinases in neuroinflammation and cerebral ischemia. Neuroscience 2009,158, 983–994. [CrossRef]
72.
Qin, W.; Li, J.; Zhu, R.; Gao, S.; Fan, J.; Xia, M.; Zhao, R.C.; Zhang, J. Melatonin protects blood-brain barrier integrity and
permeability by inhibiting matrix metalloproteinase-9 via the NOTCH3/NF-
κ
B pathway. Aging
2019
,11, 11391–11415. [CrossRef]
73.
Yabluchanskiy, A.; Ma, Y.; Iyer, R.P.; Hall, M.E.; Lindsey, M.L. Matrix Metalloproteinase-9: Many Shades of Function in
Cardiovascular Disease. Physiology 2013,28, 391–403. [CrossRef] [PubMed]
74. Nakagawa, S.; Castro, V.; Toborek, M. Infection of human pericytes by HIV-1 disrupts the integrity of the blood-brain barrier. J.
Cell. Mol. Med. 2012,16, 2950–2957. [CrossRef] [PubMed]
75.
Ralay Ranaivo, H.; Hodge, J.N.; Choi, N.; Wainwright, M.S. Albumin induces upregulation of matrix metalloproteinase-9 in
astrocytes via MAPK and reactive oxygen species-dependent pathways. J. Neuroinflammation 2012,9, 68. [CrossRef] [PubMed]
76.
Kim, Y.S.; Kim, S.S.; Cho, J.J.; Choi, D.H.; Hwang, O.; Shin, D.H.; Chun, H.S.; Beal, M.F.; Joh, T.H. Matrix metalloproteinase-3: A
novel signaling proteinase from apoptotic neuronal cells that activates microglia. J. Neurosci. 2005,25, 3701–3711. [CrossRef]
77.
Barr, T.L.; Latour, L.L.; Lee, K.Y.; Schaewe, T.J.; Luby, M.; Chang, G.S.; El-Zammar, Z.; Alam, S.; Hallenbeck, J.M.; Kidwell, C.S.;
et al. Blood-brain barrier disruption in humans is independently associated with increased matrix metalloproteinase-9. Stroke
2010,41, e123–e128. [CrossRef]
78.
Weekman, E.M.; Wilcock, D.M. Matrix Metalloproteinase in Blood-Brain Barrier Breakdown in Dementia. J. Alzheimer’s Dis. JAD
2016,49, 893–903. [CrossRef]
79.
Zhang, S.; An, Q.; Wang, T.; Gao, S.; Zhou, G. Autophagy- and MMP-2/9-mediated Reduction and Redistribution of ZO-1
Contribute to Hyperglycemia-increased Blood-Brain Barrier Permeability During Early Reperfusion in Stroke. Neuroscience
2018
,
377, 126–137. [CrossRef]
80.
Rempe, R.G.; Hartz, A.M.S.; Soldner, E.L.B.; Sokola, B.S.; Alluri, S.R.; Abner, E.L.; Kryscio, R.J.; Pekcec, A.; Schlichtiger, J.; Bauer,
B. Matrix Metalloproteinase-Mediated Blood-Brain Barrier Dysfunction in Epilepsy. J. Neurosci.
2018
,38, 4301–4315. [CrossRef]
81.
Takata, F.; Nakagawa, S.; Matsumoto, J.; Dohgu, S. Blood-Brain Barrier Dysfunction Amplifies the Development of Neuroin-
flammation: Understanding of Cellular Events in Brain Microvascular Endothelial Cells for Prevention and Treatment of BBB
Dysfunction. Front. Cell. Neurosci. 2021,15, 661838. [CrossRef]
82.
Chung, T.D.; Linville, R.M.; Guo, Z.; Ye, R.; Jha, R.; Grifno, G.N.; Searson, P.C. Effects of acute and chronic oxidative stress on the
blood-brain barrier in 2D and 3D in vitro models. Fluids Barriers CNS 2022,19, 33. [CrossRef] [PubMed]
83.
Niedzielska, E.; Smaga, I.; Gawlik, M.; Moniczewski, A.; Stankowicz, P.; Pera, J.; Filip, M. Oxidative Stress in Neurodegenerative
Diseases. Mol. Neurobiol. 2016,53, 4094–4125. [CrossRef] [PubMed]
84.
Liguori, I.; Russo, G.; Curcio, F.; Bulli, G.; Aran, L.; Della-Morte, D.; Gargiulo, G.; Testa, G.; Cacciatore, F.; Bonaduce, D.; et al.
Oxidative stress, aging, and diseases. Clin. Interv. Aging 2018,13, 757–772. [CrossRef] [PubMed]
85.
Banks, W.A.; Rhea, E.M. The Blood-Brain Barrier, Oxidative Stress, and Insulin Resistance. Antioxidants
2021
,10, 1695. [CrossRef]
86.
Carvalho, C.; Moreira, P.I. Oxidative Stress: A Major Player in Cerebrovascular Alterations Associated to Neurodegenerative
Events. Front. Physiol. 2018,9, 806. [CrossRef] [PubMed]
87.
Cameron, T.; Bennet, T.; Rowe, E.M.; Anwer, M.; Wellington, C.L.; Cheung, K.C. Review of Design Considerations for Brain-on-a-
Chip Models. Micromachines 2021,12, 441. [CrossRef] [PubMed]
88.
Gradišnik, L.; Bošnjak, R.; Bunc, G.; Ravnik, J.; Maver, T.; Velnar, T. Neurosurgical Approaches to Brain Tissue Harvesting for the
Establishment of Cell Cultures in Neural Experimental Cell Models. Materials 2021,14, 6857. [CrossRef]
89.
Tran, M.; Heo, C.; Lee, L.P.; Cho, H. Human mini-blood-brain barrier models for biomedical neuroscience research: A review.
Biomater. Res. 2022,26, 82. [CrossRef]
90.
Bhalerao, A.; Sivandzade, F.; Archie, S.R.; Chowdhury, E.A.; Noorani, B.; Cucullo, L. In vitro modeling of the neurovascular unit:
Advances in the field. Fluids Barriers CNS 2020,17, 22. [CrossRef]
91.
Eigenmann, D.E.; Xue, G.; Kim, K.S.; Moses, A.V.; Hamburger, M.; Oufir, M. Comparative study of four immortalized human
brain capillary endothelial cell lines, hCMEC/D3, hBMEC, TY10, and BB19, and optimization of culture conditions, for an in vitro
blood-brain barrier model for drug permeability studies. Fluids Barriers CNS 2013,10, 33. [CrossRef]
92.
Urich, E.; Lazic, S.E.; Molnos, J.; Wells, I.; Freskgård, P.O. Transcriptional profiling of human brain endothelial cells reveals key
properties crucial for predictive in vitro blood-brain barrier models. PLoS ONE 2012,7, e38149. [CrossRef]
93.
Delsing, L.; Herland, A.; Falk, A.; Hicks, R.; Synnergren, J.; Zetterberg, H. Models of the blood-brain barrier using iPSC-derived
cells. Mol. Cell. Neurosci. 2020,107, 103533. [CrossRef]
94.
Dong, T.; Li, M.; Gao, F.; Wei, P.; Wang, J. Construction and imaging of a neurovascular unit model. Neural Regen. Res.
2022
,
17, 1685–1694. [CrossRef]
95.
Kolagar, T.A.; Farzaneh, M.; Nikkar, N.; Khoshnam, S.E. Human Pluripotent Stem Cells in Neurodegenerative Diseases: Potentials,
Advances and Limitations. Curr. Stem Cell Res. Ther. 2020,15, 102–110. [CrossRef]
96.
Ge, S.; Jiang, X.; Paul, D.; Song, L.; Wang, X.; Pachter, J.S. Human ES-derived MSCs correct TNF-
α
-mediated alterations in a
blood-brain barrier model. Fluids Barriers CNS 2019,16, 18. [CrossRef]
97.
Tian, X.; Brookes, O.; Battaglia, G. Pericytes from Mesenchymal Stem Cells as a model for the blood-brain barrier. Sci. Rep.
2017
,
7, 39676. [CrossRef] [PubMed]
98.
Uwamori, H.; Ono, Y.; Yamashita, T.; Arai, K.; Sudo, R. Comparison of organ-specific endothelial cells in terms of microvascular
formation and endothelial barrier functions. Microvasc. Res. 2019,122, 60–70. [CrossRef] [PubMed]
Int. J. Mol. Sci. 2023,24, 12699 31 of 40
99.
Tang, G.; Liu, Y.; Zhang, Z.; Lu, Y.; Wang, Y.; Huang, J.; Li, Y.; Chen, X.; Gu, X.; Wang, Y.; et al. Mesenchymal stem cells maintain
blood-brain barrier integrity by inhibiting aquaporin-4 upregulation after cerebral ischemia. Stem Cells
2014
,32, 3150–3162.
[CrossRef]
100.
Gastfriend, B.D.; Palecek, S.P.; Shusta, E.V. Modeling the blood-brain barrier: Beyond the endothelial cells. Curr. Opin. Biomed.
Eng. 2018,5, 6–12. [CrossRef]
101.
Le Roux, G.; Jarray, R.; Guyot, A.C.; Pavoni, S.; Costa, N.; Théodoro, F.; Nassor, F.; Pruvost, A.; Tournier, N.; Kiyan, Y.; et al.
Proof-of-Concept Study of Drug Brain Permeability Between in Vivo Human Brain and an in Vitro iPSCs-Human Blood-Brain
Barrier Model. Sci. Rep. 2019,9, 16310. [CrossRef] [PubMed]
102. Page, S.; Patel, R.; Raut, S.; Al-Ahmad, A. Neurological diseases at the blood-brain barrier: Stemming new scientific paradigms
using patient-derived induced pluripotent cells. Biochim. Biophys. Acta. Mol. Basis Dis.
2020
,1866, 165358. [CrossRef] [PubMed]
103.
Lippmann, E.S.; Azarin, S.M.; Kay, J.E.; Nessler, R.A.; Wilson, H.K.; Al-Ahmad, A.; Palecek, S.P.; Shusta, E.V. Derivation of
blood-brain barrier endothelial cells from human pluripotent stem cells. Nat. Biotechnol. 2012,30, 783–791. [CrossRef]
104.
Lippmann, E.S.; Azarin, S.M.; Palecek, S.P.; Shusta, E.V. Commentary on human pluripotent stem cell-based blood-brain barrier
models. Fluids Barriers CNS 2020,17, 64. [CrossRef] [PubMed]
105.
Li, Y.; Terstappen, G.C.; Zhang, W. Differentiation of Human Induced Pluripotent Stem Cells (hiPSC) into Endothelial-Type Cells
and Establishment of an In Vitro Blood-Brain Barrier Model. Methods Mol. Biol. 2022,2454, 521–530. [CrossRef] [PubMed]
106.
Neal, E.H.; Marinelli, N.A.; Shi, Y.; McClatchey, P.M.; Balotin, K.M.; Gullett, D.R.; Hagerla, K.A.; Bowman, A.B.; Ess, K.C.; Wikswo,
J.P.; et al. A Simplified, Fully Defined Differentiation Scheme for Producing Blood-Brain Barrier Endothelial Cells from Human
iPSCs. Stem Cell Rep. 2019,12, 1380–1388. [CrossRef]
107.
Lu, T.M.; Houghton, S.; Magdeldin, T.; Durán, J.G.B.; Minotti, A.P.; Snead, A.; Sproul, A.; Nguyen, D.H.T.; Xiang, J.; Fine, H.A.;
et al. Pluripotent stem cell-derived epithelium misidentified as brain microvascular endothelium requires ETS factors to acquire
vascular fate. Proc. Natl. Acad. Sci. USA 2021,118, e0118. [CrossRef]
108.
Lim, R.G.; Quan, C.; Reyes-Ortiz, A.M.; Lutz, S.E.; Kedaigle, A.J.; Gipson, T.A.; Wu, J.; Vatine, G.D.; Stocksdale, J.; Casale,
M.S.; et al. Huntington’s Disease iPSC-Derived Brain Microvascular Endothelial Cells Reveal WNT-Mediated Angiogenic and
Blood-Brain Barrier Deficits. Cell Rep. 2017,19, 1365–1377. [CrossRef]
109.
Vignone, D.; Gonzalez Paz, O.; Fini, I.; Cellucci, A.; Auciello, G.; Battista, M.R.; Gloaguen, I.; Fortuni, S.; Cariulo, C.; Khetarpal, V.;
et al. Modelling the Human Blood-Brain Barrier in Huntington Disease. Int. J. Mol. Sci. 2022,23, 7813. [CrossRef]
110.
Andjelkovic, A.V.; Stamatovic, S.M.; Phillips, C.M.; Martinez-Revollar, G.; Keep, R.F. Modeling blood-brain barrier pathology in
cerebrovascular disease in vitro: Current and future paradigms. Fluids Barriers CNS 2020,17, 44. [CrossRef]
111.
Villabona-Rueda, A.; Erice, C.; Pardo, C.A.; Stins, M.F. The Evolving Concept of the Blood Brain Barrier (BBB): From a Single
Static Barrier to a Heterogeneous and Dynamic Relay Center. Front. Cell. Neurosci. 2019,13, 405. [CrossRef]
112.
Noumbissi, M.E.; Galasso, B.; Stins, M.F. Brain vascular heterogeneity: Implications for disease pathogenesis and design of
in vitro blood-brain barrier models. Fluids Barriers CNS 2018,15, 12. [CrossRef] [PubMed]
113.
Vera, D.; García-Díaz, M.; Torras, N.; Álvarez, M.; Villa, R.; Martinez, E. Engineering Tissue Barrier Models on Hydrogel
Microfluidic Platforms. ACS Appl. Mater. Interfaces 2021,13, 13920–13933. [CrossRef] [PubMed]
114.
Huang, G.; Li, F.; Zhao, X.; Ma, Y.; Li, Y.; Lin, M.; Jin, G.; Lu, T.J.; Genin, G.M.; Xu, F. Functional and Biomimetic Materials for
Engineering of the Three-Dimensional Cell Microenvironment. Chem. Rev. 2017,117, 12764–12850. [CrossRef] [PubMed]
115.
Kleinman, H.K.; Kim, K.; Kang, H. Matrigel uses in cell biology and for the identification of thymosin
β
4, a mediator of tissue
regeneration. Appl. Biol. Chem. 2018,61, 703–708. [CrossRef]
116.
Draget, K.I.; Taylor, C. Chemical, physical and biological properties of alginates and their biomedical implications. Food Hydrocoll.
2011,25, 251–256. [CrossRef]
117.
Catoira, M.C.; Fusaro, L.; Di Francesco, D.; Ramella, M.; Boccafoschi, F. Overview of natural hydrogels for regenerative medicine
applications. J. Mater. Sci. Mater. Med. 2019,30, 115. [CrossRef]
118.
Vasile, C.; Pamfil, D.; Stoleru, E.; Baican, M. New Developments in Medical Applications of Hybrid Hydrogels Containing
Natural Polymers. Molecules 2020,25, 1539. [CrossRef]
119.
Samanipour, R.; Tahmooressi, H.; Rezaei Nejad, H.; Hirano, M.; Shin, S.R.; Hoorfar, M. A review on 3D printing functional brain
model. Biomicrofluidics 2022,16, 011501. [CrossRef]
120.
GROBSTEIN, C. Morphogenetic interaction between embryonic mouse tissues separated by a membrane filter. Nature
1953
,
172, 869–870. [CrossRef]
121.
Raut, S.; Bhalerao, A.; Noorani, B.; Cucullo, L. In Vitro Models of the Blood-Brain Barrier. Methods Mol. Biol.
2022
,2492, 25–49.
[CrossRef]
122.
Stone, N.L.; England, T.J.; O’Sullivan, S.E. A Novel Transwell Blood Brain Barrier Model Using Primary Human Cells. Front. Cell.
Neurosci. 2019,13, 230. [CrossRef] [PubMed]
123.
Nakayama-Kitamura, K.; Shigemoto-Mogami, Y.; Toyoda, H.; Mihara, I.; Moriguchi, H.; Naraoka, H.; Furihata, T.; Ishida, S.;
Sato, K. Usefulness of a humanized tricellular static transwell blood-brain barrier model as a microphysiological system for drug
development applications—A case study based on the benchmark evaluations of blood-brain barrier microphysiological system.
Regen. Ther. 2023,22, 192–202. [CrossRef] [PubMed]
124.
Kramer, N.; Walzl, A.; Unger, C.; Rosner, M.; Krupitza, G.; Hengstschläger, M.; Dolznig, H. In vitro cell migration and invasion
assays. Mutat. Res. 2013,752, 10–24. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 32 of 40
125.
Park, J.S.; Choe, K.; Khan, A.; Jo, M.H.; Park, H.Y.; Kang, M.H.; Park, T.J.; Kim, M.O. Establishing Co-Culture Blood-Brain Barrier
Models for Different Neurodegeneration Conditions to Understand Its Effect on BBB Integrity. Int. J. Mol. Sci.
2023
,24, 5283.
[CrossRef]
126.
Appelt-Menzel, A.; Cubukova, A.; Günther, K.; Edenhofer, F.; Piontek, J.; Krause, G.; Stüber, T.; Walles, H.; Neuhaus, W.; Metzger,
M. Establishment of a Human Blood-Brain Barrier Co-culture Model Mimicking the Neurovascular Unit Using Induced Pluri-
and Multipotent Stem Cells. Stem Cell Rep. 2017,8, 894–906. [CrossRef]
127.
Hollmann, E.K.; Bailey, A.K.; Potharazu, A.V.; Neely, M.D.; Bowman, A.B.; Lippmann, E.S. Accelerated differentiation of human
induced pluripotent stem cells to blood-brain barrier endothelial cells. Fluids Barriers CNS 2017,14, 9. [CrossRef]
128.
Lippmann, E.S.; Al-Ahmad, A.; Azarin, S.M.; Palecek, S.P.; Shusta, E.V. A retinoic acid-enhanced, multicellular human blood-brain
barrier model derived from stem cell sources. Sci. Rep. 2014,4, 4160. [CrossRef] [PubMed]
129.
Singh, N.R.; Gromnicova, R.; Brachner, A.; Kraev, I.; Romero, I.A.; Neuhaus, W.; Male, D. A hydrogel model of the human
blood-brain barrier using differentiated stem cells. PLoS ONE 2023,18, e0283954. [CrossRef]
130.
Hawkins, B.T.; Grego, S.; Sellgren, K.L. Three-dimensional culture conditions differentially affect astrocyte modulation of brain
endothelial barrier function in response to transforming growth factor β1. Brain Res. 2015,1608, 167–176. [CrossRef]
131.
Kim, W.; Kim, J.; Lee, S.Y.; Kim, H.M.; Joo, K.M.; Nam, D.H. Simplified in vitro 3D co-culture-based blood-brain barrier model
using transwell. Biochem. Biophys. Res. Commun. 2022,620, 63–68. [CrossRef]
132.
Zakharova, M.; Tibbe, M.P.; Koch, L.S.; Le-The, H.; Leferink, A.M.; den Berg, A.; Meer, A.D.; Broersen, K.; Segerink, L.I. Transwell–
Integrated 2
µ
m Thick Transparent Polydimethylsiloxane Membranes with Controlled Pore Sizes and Distribution to Model the
Blood–Brain Barrier. Adv. Mater. Technol. 2021,6, 2100138. [CrossRef]
133.
Herland, A.; van der Meer, A.D.; FitzGerald, E.A.; Park, T.E.; Sleeboom, J.J.F.; Ingber, D.E. Distinct Contributions of Astrocytes
and Pericytes to Neuroinflammation Identified in a 3D Human Blood-Brain Barrier on a Chip. PLoS ONE
2016
,11, e0150360.
[CrossRef]
134.
Cucullo, L.; Hossain, M.; Puvenna, V.; Marchi, N.; Janigro, D. The role of shear stress in Blood-Brain Barrier endothelial physiology.
BMC Neurosci. 2011,12, 40. [CrossRef]
135.
Kawakita, S.; Mandal, K.; Mou, L.; Mecwan, M.M.; Zhu, Y.; Li, S.; Sharma, S.; Hernandez, A.L.; Nguyen, H.T.; Maity, S.; et al.
Organ-On-A-Chip Models of the Blood-Brain Barrier: Recent Advances and Future Prospects. Small
2022
,18, e2201401. [CrossRef]
136.
Nader, E.; Skinner, S.; Romana, M.; Fort, R.; Lemonne, N.; Guillot, N.; Gauthier, A.; Antoine-Jonville, S.; Renoux, C.; Hardy-
Dessources, M.D.; et al. Blood Rheology: Key Parameters, Impact on Blood Flow, Role in Sickle Cell Disease and Effects of
Exercise. Front. Physiol. 2019,10, 1329. [CrossRef]
137.
Bi, W.; Cai, S.; Lei, T.; Wang, L. Implementation of blood-brain barrier on microfluidic chip: Recent advance and future prospects.
Ageing Res. Rev. 2023,87, 101921. [CrossRef]
138.
Oddo, A.; Peng, B.; Tong, Z.; Wei, Y.; Tong, W.Y.; Thissen, H.; Voelcker, N.H. Advances in Microfluidic Blood-Brain Barrier (BBB)
Models. Trends Biotechnol. 2019,37, 1295–1314. [CrossRef]
139.
Byun, C.K.; Abi-Samra, K.; Cho, Y.K.; Takayama, S. Pumps for microfluidic cell culture. Electrophoresis
2014
,35, 245–257.
[CrossRef]
140.
Iakovlev, A.P.; Erofeev, A.S.; Gorelkin, P.V. Novel Pumping Methods for Microfluidic Devices: A Comprehensive Review.
Biosensors 2022,12, 956. [CrossRef]
141.
Garcia-Polite, F.; Martorell, J.; Del Rey-Puech, P.; Melgar-Lesmes, P.; O’Brien, C.C.; Roquer, J.; Ois, A.; Principe, A.; Edelman, E.R.;
Balcells, M. Pulsatility and high shear stress deteriorate barrier phenotype in brain microvascular endothelium. J. Cereb. Blood
Flow Metab. 2017,37, 2614–2625. [CrossRef]
142.
Faley, S.L.; Neal, E.H.; Wang, J.X.; Bosworth, A.M.; Weber, C.M.; Balotin, K.M.; Lippmann, E.S.; Bellan, L.M. iPSC-Derived Brain
Endothelium Exhibits Stable, Long-Term Barrier Function in Perfused Hydrogel Scaffolds. Stem Cell Rep.
2019
,12, 474–487.
[CrossRef] [PubMed]
143.
Yang, M.; Sun, N.; Luo, Y.; Lai, X.; Li, P.; Zhang, Z. Emergence of debubblers in microfluidics: A critical review. Biomicrofluidics
2022,16, 031503. [CrossRef] [PubMed]
144.
Stanness, K.A.; Westrum, L.E.; Fornaciari, E.; Mascagni, P.; Nelson, J.A.; Stenglein, S.G.; Myers, T.; Janigro, D. Morphological and
functional characterization of an in vitro blood-brain barrier model. Brain Res. 1997,771, 329–342. [CrossRef] [PubMed]
145.
Janigro, D.; Leaman, S.M.; Stanness, K.A. Dynamic modeling of the blood-brain barrier: A novel tool for studies of drug delivery
to the brain. Pharm. Sci. Technol. Today 1999,2, 7–12. [CrossRef] [PubMed]
146.
Gökaltun, A.; Kang, Y.B.A.; Yarmush, M.L.; Usta, O.B.; Asatekin, A. Simple Surface Modification of Poly(dimethylsiloxane) via
Surface Segregating Smart Polymers for Biomicrofluidics. Sci. Rep. 2019,9, 7377. [CrossRef]
147.
Cho, H.; Seo, J.H.; Wong, K.H.K.; Terasaki, Y.; Park, J.; Bong, K.; Arai, K.; Lo, E.H.; Irimia, D. Three-Dimensional Blood-Brain
Barrier Model for in vitro Studies of Neurovascular Pathology. Sci. Rep. 2015,5, 15222. [CrossRef]
148.
Partyka, P.P.; Godsey, G.A.; Galie, J.R.; Kosciuk, M.C.; Acharya, N.K.; Nagele, R.G.; Galie, P.A. Mechanical stress regulates
transport in a compliant 3D model of the blood-brain barrier. Biomaterials 2017,115, 30–39. [CrossRef]
149.
Galpayage Dona, K.N.U.; Hale, J.F.; Salako, T.; Anandanatarajan, A.; Tran, K.A.; DeOre, B.J.; Galie, P.A.; Ramirez, S.H.; Andrews,
A.M. The Use of Tissue Engineering to Fabricate Perfusable 3D Brain Microvessels in vitro. Front. Physiol.
2021
,12, 715431.
[CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 33 of 40
150.
Campisi, M.; Shin, Y.; Osaki, T.; Hajal, C.; Chiono, V.; Kamm, R.D. 3D self-organized microvascular model of the human
blood-brain barrier with endothelial cells, pericytes and astrocytes. Biomaterials 2018,180, 117–129. [CrossRef] [PubMed]
151.
Nikolakopoulou, P.; Rauti, R.; Voulgaris, D.; Shlomy, I.; Maoz, B.M.; Herland, A. Recent progress in translational engineered
in vitro models of the central nervous system. Brain J. Neurol. 2020,143, 3181–3213. [CrossRef]
152.
Jagtiani, E.; Yeolekar, M.; Naik, S.; Patravale, V. In vitro blood brain barrier models: An overview. J. Control. Release
2022
,
343, 13–30. [CrossRef] [PubMed]
153.
Xu, T.; Jin, J.; Gregory, C.; Hickman, J.J.J.J.; Boland, T. Inkjet printing of viable mammalian cells. Biomaterials
2005
,26, 93–99.
[CrossRef]
154.
Mandrycky, C.; Wang, Z.; Kim, K.; Kim, D.H. 3D bioprinting for engineering complex tissues. Biotechnol. Adv.
2016
,34, 422–434.
[CrossRef]
155.
Khalil, S.; Sun, W. Biopolymer deposition for freeform fabrication of hydrogel tissue constructs. Mater. Sci. Eng. C
2007
,
27, 469–478. [CrossRef]
156.
Marino, A.; Tricinci, O.; Battaglini, M.; Filippeschi, C.; Mattoli, V.; Sinibaldi, E.; Ciofani, G. A 3D Real-Scale, Biomimetic, and
Biohybrid Model of the Blood-Brain Barrier Fabricated through Two-Photon Lithography. Small 2018,14, 2959. [CrossRef]
157.
Tricinci, O.; de Pasquale, D.; Marino, A.; Battaglini, M.; Pucci, C.; Ciofani, G. A 3D Biohybrid Real-Scale Model of the Brain
Cancer Microenvironment for Advanced In Vitro Testing. Adv. Mater. Technol. 2020,5, 2000540. [CrossRef]
158.
Kim, J.A.; Kim, H.N.; Im, S.K.; Chung, S.; Kang, J.Y.; Choi, N. Collagen-based brain microvasculature model in vitro using
three-dimensional printed template. Biomicrofluidics 2015,9, 7508. [CrossRef]
159.
Wang, Y.I.; Abaci, H.E.; Shuler, M.L. Microfluidic blood-brain barrier model provides in vivo-like barrier properties for drug
permeability screening. Biotechnol. Bioeng. 2017,114, 184–194. [CrossRef]
160.
Gunti, S.; Hoke, A.T.; Vu, K.P.; London, N.R. Organoid and Spheroid Tumor Models: Techniques and Applications. Cancers
2021
,
13, 874. [CrossRef]
161.
Lancaster, M.A.; Renner, M.; Martin, C.A.; Wenzel, D.; Bicknell, L.S.; Hurles, M.E.; Homfray, T.; Penninger, J.M.; Jackson, A.P.;
Knoblich, J.A. Cerebral organoids model human brain development and microcephaly. Nature 2013,501, 373–379. [CrossRef]
162.
Smits, L.M.; Schwamborn, J.C. Midbrain Organoids: A New Tool to Investigate Parkinson’s Disease. Front. Cell Dev. Biol.
2020
,
8, 359. [CrossRef] [PubMed]
163.
Nzou, G.; Wicks, R.T.; Wicks, E.E.; Seale, S.A.; Sane, C.H.; Chen, A.; Murphy, S.V.; Jackson, J.D.; Atala, A.J. Human Cortex
Spheroid with a Functional Blood Brain Barrier for High-Throughput Neurotoxicity Screening and Disease Modeling. Sci. Rep.
2018,8, 7413. [CrossRef] [PubMed]
164.
de Kleijn, K.M.A.; Zuure, W.A.; Straasheijm, K.R.; Martens, M.B.; Avramut, M.C.; Koning, R.I.; Martens, G.J.M. Human cortical
spheroids with a high diversity of innately developing brain cell types. Stem Cell Res. Ther. 2023,14, 50. [CrossRef] [PubMed]
165.
Wörsdörfer, P.; Dalda, N.; Kern, A.; Krüger, S.; Wagner, N.; Kwok, C.K.; Henke, E.; Ergün, S. Generation of complex human
organoid models including vascular networks by incorporation of mesodermal progenitor cells. Sci. Rep.
2019
,9, 15663.
[CrossRef]
166.
Ham, O.; Jin, Y.B.; Kim, J.; Lee, M.O. Blood vessel formation in cerebral organoids formed from human embryonic stem cells.
Biochem. Biophys. Res. Commun. 2020,521, 84–90. [CrossRef]
167.
Hol, E.M.; Pekny, M. Glial fibrillary acidic protein (GFAP) and the astrocyte intermediate filament system in diseases of the
central nervous system. Curr. Opin. Cell Biol. 2015,32, 121–130. [CrossRef]
168.
Mader, S.; Brimberg, L. Aquaporin-4 Water Channel in the Brain and Its Implication for Health and Disease. Cells
2019
,8, 90.
[CrossRef]
169.
Smyth, L.C.D.; Rustenhoven, J.; Scotter, E.L.; Schweder, P.; Faull, R.L.M.; Park, T.I.H.; Dragunow, M. Markers for human brain
pericytes and smooth muscle cells. J. Chem. Neuroanat. 2018,92, 48–60. [CrossRef]
170.
Michalski, D.; Keck, A.L.; Grosche, J.; Martens, H.; Härtig, W. Immunosignals of Oligodendrocyte Markers and Myelin-Associated
Proteins Are Critically Affected after Experimental Stroke in Wild-Type and Alzheimer Modeling Mice of Different Ages. Front.
Cell. Neurosci. 2018,12, 23. [CrossRef]
171.
Jurga, A.M.; Paleczna, M.; Kuter, K.Z. Overview of General and Discriminating Markers of Differential Microglia Phenotypes.
Front. Cell. Neurosci. 2020,14, 198. [CrossRef]
172.
Bernal, A.; Arranz, L. Nestin-expressing progenitor cells: Function, identity and therapeutic implications. Cell. Mol. Life Sci.
CMLS 2018,75, 2177–2195. [CrossRef] [PubMed]
173.
Bolte, S.; Cordelières, F.P. A guided tour into subcellular colocalization analysis in light microscopy. J. Microsc.
2006
,224, 213–232.
[CrossRef] [PubMed]
174.
Dominijanni, A.J.; Devarasetty, M.; Forsythe, S.D.; Votanopoulos, K.I.; Soker, S. Cell Viability Assays in Three-Dimensional
Hydrogels: A Comparative Study of Accuracy. Tissue Eng. Part C Methods 2021,27, 401–410. [CrossRef] [PubMed]
175.
Deng, H.; Yue, J.K.; Zusman, B.E.; Nwachuku, E.L.; Abou-Al-Shaar, H.; Upadhyayula, P.S.; Okonkwo, D.O.; Puccio, A.M. B-Cell
Lymphoma 2 (Bcl-2) and Regulation of Apoptosis after Traumatic Brain Injury: A Clinical Perspective. Medicina
2020
,56, 300.
[CrossRef]
176.
Ng Kee Kwong, K.C.; Gregory, J.M.; Pal, S.; Chandran, S.; Mehta, A.R. Cerebrospinal fluid cytotoxicity in amyotrophic lateral
sclerosis: A systematic review of in vitro studies. Brain Commun. 2020,2, fcaa121. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 34 of 40
177.
Butt, A.M.; Jones, H.C.; Abbott, N.J. Electrical resistance across the blood-brain barrier in anaesthetized rats: A developmental
study. J. Physiol. 1990,429, 47–62. [CrossRef]
178. Crone, C.; Olesen, S.P. Electrical resistance of brain microvascular endothelium. Brain Res. 1982,241, 49–55. [CrossRef]
179.
Jeong, S.; Kim, S.; Buonocore, J.; Park, J.; Welsh, C.J.; Li, J.; Han, A. A Three-Dimensional Arrayed Microfluidic Blood-Brain
Barrier Model With Integrated Electrical Sensor Array. IEEE Trans. Bio-Med Eng. 2018,65, 431–439. [CrossRef]
180.
Park, T.E.; Mustafaoglu, N.; Herland, A.; Hasselkus, R.; Mannix, R.; FitzGerald, E.A.; Prantil-Baun, R.; Watters, A.; Henry, O.;
Benz, M.; et al. Hypoxia-enhanced Blood-Brain Barrier Chip recapitulates human barrier function and shuttling of drugs and
antibodies. Nat. Commun. 2019,10, 2621. [CrossRef]
181.
Vigh, J.P.; Kincses, A.; Ozgür, B.; Walter, F.R.; Santa-Maria, A.R.; Valkai, S.; Vastag, M.; Neuhaus, W.; Brodin, B.; Dér, A.; et al.
Transendothelial Electrical Resistance Measurement across the Blood-Brain Barrier: A Critical Review of Methods. Micromachines
2021,12, 685. [CrossRef]
182.
Palma-Florez, S.; López-Canosa, A.; Moralez-Zavala, F.; Castaño, O.; Kogan, M.J.; Samitier, J.; Lagunas, A.; Mir, M. BBB-on-a-chip
with integrated micro-TEER for permeability evaluation of multi-functionalized gold nanorods against Alzheimer’s disease.
J. Nanobiotechnol. 2023,21, 115. [CrossRef] [PubMed]
183.
Jeftha, T.; Makhathini, K.B.; Fisher, D. The Effect of dexamethasone on lipopolysaccharide-induced Inflammation of Endothelial
Cells of the Blood-brain Barrier/Brain Capillaries. Curr. Neurovasc. Res. 2023, ahead of print. [CrossRef] [PubMed]
184.
Förster, C.; Burek, M.; Romero, I.A.; Weksler, B.; Couraud, P.O.; Drenckhahn, D. Differential effects of hydrocortisone and TNF
α
on tight junction proteins in an in vitro model of the human blood–brain barrier. J. Physiol.
2008
,586, 1937–1949. [CrossRef]
[PubMed]
185.
Hajal, C.; Offeddu, G.S.; Shin, Y.; Zhang, S.; Morozova, O.; Hickman, D.; Knutson, C.G.; Kamm, R.D. Engineered human
blood-brain barrier microfluidic model for vascular permeability analyses. Nat. Protoc. 2022,17, 95–128. [CrossRef]
186.
Wong, A.D.; Ye, M.; Levy, A.F.; Rothstein, J.D.; Bergles, D.E.; Searson, P.C. The blood-brain barrier: An engineering perspective.
Front. Neuroeng. 2013,6, 7. [CrossRef]
187.
Helms, H.C.; Abbott, N.J.; Burek, M.; Cecchelli, R.; Couraud, P.O.; Deli, M.A.; Förster, C.; Galla, H.J.; Romero, I.A.; Shusta, E.V.;
et al. In vitro models of the blood-brain barrier: An overview of commonly used brain endothelial cell culture models and
guidelines for their use. J. Cereb. Blood Flow Metab. 2016,36, 862–890. [CrossRef]
188.
Brown, R.C.; Lockwood, A.H.; Sonawane, B.R. Neurodegenerative diseases: An overview of environmental risk factors. Environ.
Health Perspect. 2005,113, 1250–1256. [CrossRef]
189. Prüss, H. Autoantibodies in neurological disease. Nat. Rev. Immunol. 2021,21, 798–813. [CrossRef]
190.
Ramanathan, S.; Brilot, F.; Irani, S.R.; Dale, R.C. Origins and immunopathogenesis of autoimmune central nervous system
disorders. Nat. Rev. Neurol. 2023,19, 172–190. [CrossRef]
191. Lassmann, H. Multiple Sclerosis Pathology. Cold Spring Harb. Perspect. Med. 2018,8, a028936. [CrossRef]
192.
Sun, Y.; Yu, H.; Guan, Y. Glia Connect Inflammation and Neurodegeneration in Multiple Sclerosis. Neurosci. Bull.
2023
,
39, 466–478. [CrossRef]
193.
Angelini, G.; Bani, A.; Constantin, G.; Rossi, B. The interplay between T helper cells and brain barriers in the pathogenesis of
multiple sclerosis. Front. Cell. Neurosci. 2023,17, 1101379. [CrossRef] [PubMed]
194.
Li, L.Y.; Kreye, J.; Burek, M.; Cordero-Gomez, C.; Barthel, P.C.; Sánchez-Sendín, E.; Kornau, H.C.; Schmitz, D.; Scharf, M.;
Meybohm, P.; et al. Brain blood vessel autoantibodies in patients with NMDA and GABAA receptor encephalitis: Identification
of unconventional Myosin-X as target antigen. Front. Cell. Neurosci. 2023,17, 1077204. [CrossRef] [PubMed]
195.
Lerch, M.; Schanda, K.; Lafon, E.; Würzner, R.; Mariotto, S.; Dinoto, A.; Wendel, E.M.; Lechner, C.; Hegen, H.; Rostásy, K.;
et al. More Efficient Complement Activation by Anti-Aquaporin-4 Compared With Anti-Myelin Oligodendrocyte Glycoprotein
Antibodies. Neurol. Neuroimmunol. Neuroinflamm. 2023,10, 59. [CrossRef] [PubMed]
196.
Jarius, S.; Aktas, O.; Ayzenberg, I.; Bellmann-Strobl, J.; Berthele, A.; Giglhuber, K.; Häußler, V.; Havla, J.; Hellwig, K.; Hümmert,
M.W.; et al. Update on the diagnosis and treatment of neuromyelits optica spectrum disorders (NMOSD)—Revised recommen-
dations of the Neuromyelitis Optica Study Group (NEMOS). Part I: Diagnosis and differential diagnosis. J. Neurol.
2023
,270,
3341–3368. [CrossRef]
197.
Banwell, B.; Bennett, J.L.; Marignier, R.; Kim, H.J.; Brilot, F.; Flanagan, E.P.; Ramanathan, S.; Waters, P.; Tenembaum, S.; Graves,
J.S.; et al. Diagnosis of myelin oligodendrocyte glycoprotein antibody-associated disease: International MOGAD Panel proposed
criteria. Lancet Neurol. 2023,22, 268–282. [CrossRef] [PubMed]
198. Boveri, M.; Kinsner, A.; Berezowski, V.; Lenfant, A.M.; Draing, C.; Cecchelli, R.; Dehouck, M.P.; Hartung, T.; Prieto, P.; Bal-Price,
A. Highly purified lipoteichoic acid from gram-positive bacteria induces in vitro blood-brain barrier disruption through glia
activation: Role of pro-inflammatory cytokines and nitric oxide. Neuroscience 2006,137, 1193–1209. [CrossRef]
199. Brown, G.C. The endotoxin hypothesis of neurodegeneration. J. Neuroinflamm. 2019,16, 180. [CrossRef]
200.
Simöes Da Gama, C.; Morin-Brureau, M. Study of BBB Dysregulation in Neuropathogenicity Using Integrative Human Model of
Blood-Brain Barrier. Front. Cell. Neurosci. 2022,16, 863836. [CrossRef]
201.
Ding, G.; Shao, Q.; Yu, H.; Liu, J.; Li, Y.; Wang, B.; Sang, H.; Li, D.; Bing, A.; Hou, Y.; et al. Tight Junctions, the Key Factor in
Virus-Related Disease. Pathogens 2022,11, 1200. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 35 of 40
202.
Voirin, A.C.; Chatard, M.; Briançon-Marjollet, A.; Pepin, J.L.; Perek, N.; Roche, F. Loss of Blood-Brain Barrier Integrity in an
In Vitro Model Subjected to Intermittent Hypoxia: Is Reversion Possible with a HIF-1αPathway Inhibitor? Int. J. Mol. Sci. 2023,
24, 5062. [CrossRef]
203.
Archie, S.R.; Al Shoyaib, A.; Cucullo, L. Blood-Brain Barrier Dysfunction in CNS Disorders and Putative Therapeutic Targets: An
Overview. Pharmaceutics 2021,13, 1779. [CrossRef] [PubMed]
204.
Carlsson, R.; Enström, A.; Paul, G. Molecular Regulation of the Response of Brain Pericytes to Hypoxia. Int. J. Mol. Sci.
2023
,24,
5671. [CrossRef] [PubMed]
205.
Monroy, G.R.; Murguiondo Pérez, R.; Weintraub Ben Zión, E.; Vidal Alcántar-Garibay, O.; Loza-López, E.C.; Tejerina Marion,
E.; Blancarte Hernández, E.; Navarro-Torres, L.; Ibarra, A. Immunization with Neural-Derived Peptides in Neurodegenerative
Diseases: A Narrative Review. Biomedicines 2023,11, 919. [CrossRef] [PubMed]
206.
Simon, D.W.; McGeachy, M.J.; Bayır, H.; Clark, R.S.B.; Loane, D.J.; Kochanek, P.M. The far-reaching scope of neuroinflammation
after traumatic brain injury. Nat. Rev. Neurol. 2017,13, 171–191. [CrossRef]
207.
Obermeier, B.; Daneman, R.; Ransohoff, R.M. Development, maintenance and disruption of the blood-brain barrier. Nat. Med.
2013,19, 1584–1596. [CrossRef]
208. Lapointe, S.; Perry, A.; Butowski, N.A. Primary brain tumours in adults. Lancet 2018,392, 432–446. [CrossRef] [PubMed]
209.
Anwar, M.M.; Özkan, E.; Gürsoy-Özdemir, Y. The role of extracellular matrix alterations in mediating astrocyte damage and
pericyte dysfunction in Alzheimer’s disease: A comprehensive review. Eur. J. Neurosci. 2022,56, 5453–5475. [CrossRef]
210.
Neumann, M.; Sampathu, D.M.; Kwong, L.K.; Truax, A.C.; Micsenyi, M.C.; Chou, T.T.; Bruce, J.; Schuck, T.; Grossman, M.;
Clark, C.M.; et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science
2006
,
314, 130–133. [CrossRef]
211.
Prasad, A.; Bharathi, V.; Sivalingam, V.; Girdhar, A.; Patel, B.K. Molecular Mechanisms of TDP-43 Misfolding and Pathology in
Amyotrophic Lateral Sclerosis. Front. Mol. Neurosci. 2019,12, 25. [CrossRef]
212.
Steinruecke, M.; Lonergan, R.M.; Selvaraj, B.T.; Chandran, S.; Diaz-Castro, B.; Stavrou, M. Blood-CNS barrier dysfunction
in amyotrophic lateral sclerosis: Proposed mechanisms and clinical implications. J. Cereb. Blood Flow Metab.
2023
, 1153281.
[CrossRef] [PubMed]
213.
Ghatak, S.; Trudler, D.; Dolatabadi, N.; Ambasudhan, R. Parkinson’s disease: What the model systems have taught us so far. J.
Genet. 2018,97, 729–751. [CrossRef]
214.
Moon, H.E.; Paek, S.H. Mitochondrial Dysfunction in Parkinson’s Disease. Exp. Neurobiol.
2015
,24, 103–116. [CrossRef] [PubMed]
215.
Anderson, J.P.; Walker, D.E.; Goldstein, J.M.; de Laat, R.; Banducci, K.; Caccavello, R.J.; Barbour, R.; Huang, J.; Kling, K.; Lee, M.;
et al. Phosphorylation of Ser-129 is the dominant pathological modification of alpha-synuclein in familial and sporadic Lewy
body disease. J. Biol. Chem. 2006,281, 29739–29752. [CrossRef] [PubMed]
216.
Sonustun, B.; Altay, M.F.; Strand, C.; Ebanks, K.; Hondhamuni, G.; Warner, T.T.; Lashuel, H.A.; Bandopadhyay, R. Pathological
Relevance of Post-Translationally Modified Alpha-Synuclein (pSer87, pSer129, nTyr39) in Idiopathic Parkinson’s Disease and
Multiple System Atrophy. Cells 2022,11, 906. [CrossRef] [PubMed]
217.
Bates, G.P.; Dorsey, R.; Gusella, J.F.; Hayden, M.R.; Kay, C.; Leavitt, B.R.; Nance, M.; Ross, C.A.; Scahill, R.I.; Wetzel, R.; et al.
Huntington disease. Nat. Rev. Dis. Prim. 2015,1, 15005. [CrossRef]
218.
Saba, J.; Couselo, F.L.; Bruno, J.; Carniglia, L.; Durand, D.; Lasaga, M.; Caruso, C. Neuroinflammation in Huntington’s Disease: A
Starring Role for Astrocyte and Microglia. Curr. Neuropharmacol. 2022,20, 1116–1143. [CrossRef]
219.
Diaz-Castro, B.; Gangwani, M.R.; Yu, X.; Coppola, G.; Khakh, B.S. Astrocyte molecular signatures in Huntington’s disease. Sci.
Transl. Med. 2019,11, aaw8546. [CrossRef]
220.
Balasa, R.; Barcutean, L.; Mosora, O.; Manu, D. Reviewing the Significance of Blood-Brain Barrier Disruption in Multiple Sclerosis
Pathology and Treatment. Int. J. Mol. Sci. 2021,22, 8370. [CrossRef] [PubMed]
221.
Jasiak-Zato´nska, M.; Pietrzak, A.; Wyciszkiewicz, A.; Wi˛esik-Szewczyk, E.; Pawlak-Bu´s, K.; Leszczy ´nski, P.; Kozubski, W.;
Michalak, S.; Kalinowska-Łyszczarz, A. Different blood-brain-barrier disruption profiles in multiple sclerosis, neuromyelitis
optica spectrum disorders, and neuropsychiatric systemic lupus erythematosus. Neurol. I Neurochir. Pol.
2022
,56, 246–255.
[CrossRef]
222.
Ortiz, G.G.; Pacheco-Moisés, F.P.; Macías-Islas, M.Á.; Flores-Alvarado, L.J.; Mireles-Ramírez, M.A.; González-Renovato, E.D.;
Hernández-Navarro, V.E.; Sánchez-López, A.L.; Alatorre-Jiménez, M.A. Role of the blood-brain barrier in multiple sclerosis.
Arch. Med Res. 2014,45, 687–697. [CrossRef]
223.
Spencer, J.I.; Bell, J.S.; DeLuca, G.C. Vascular pathology in multiple sclerosis: Reframing pathogenesis around the blood-brain
barrier. J. Neurol. Neurosurg. Psychiatry 2018,89, 42–52. [CrossRef]
224.
Nishihara, H.; Soldati, S.; Mossu, A.; Rosito, M.; Rudolph, H.; Muller, W.A.; Latorre, D.; Sallusto, F.; Sospedra, M.; Martin, R.; et al.
Human CD4+ T cell subsets differ in their abilities to cross endothelial and epithelial brain barriers in vitro. Fluids Barriers CNS
2020,17, 3. [CrossRef]
225.
Lengfeld, J.E.; Lutz, S.E.; Smith, J.R.; Diaconu, C.; Scott, C.; Kofman, S.B.; Choi, C.; Walsh, C.M.; Raine, C.S.; Agalliu, I.; et al.
Endothelial Wnt/
β
-catenin signaling reduces immune cell infiltration in multiple sclerosis. Proc. Natl. Acad. Sci. USA
2017
,
114, E1168–E1177. [CrossRef]
226.
Liu, J.; Xiao, Q.; Xiao, J.; Niu, C.; Li, Y.; Zhang, X.; Zhou, Z.; Shu, G.; Yin, G. Wnt/
β
-catenin signalling: Function, biological
mechanisms, and therapeutic opportunities. Signal Transduct. Target. Ther. 2022,7, 3. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 36 of 40
227.
Niu, J.; Tsai, H.H.; Hoi, K.K.; Huang, N.; Yu, G.; Kim, K.; Baranzini, S.E.; Xiao, L.; Chan, J.R.; Fancy, S.P.J. Aberrant oligodendroglial-
vascular interactions disrupt the blood-brain barrier, triggering CNS inflammation. Nat. Neurosci.
2019
,22, 709–718. [CrossRef]
[PubMed]
228.
Derada Troletti, C.; Fontijn, R.D.; Gowing, E.; Charabati, M.; van Het Hof, B.; Didouh, I.; van der Pol, S.M.A.; Geerts, D.; Prat, A.;
van Horssen, J.; et al. Inflammation-induced endothelial to mesenchymal transition promotes brain endothelial cell dysfunction
and occurs during multiple sclerosis pathophysiology. Cell Death Dis. 2019,10, 45. [CrossRef]
229.
Luo, Y.; Yang, H.; Wan, Y.; Yang, S.; Wu, J.; Chen, S.; Li, Y.; Jin, H.; He, Q.; Zhu, D.Y.; et al. Endothelial ETS1 inhibition exacerbate
blood-brain barrier dysfunction in multiple sclerosis through inducing endothelial-to-mesenchymal transition. Cell Death Dis.
2022,13, 462. [CrossRef] [PubMed]
230.
Cerutti, C.; Romero, I.A. An In Vitro Blood-Brain Barrier Model to Study Firm Shear Stress-Resistant Leukocyte Adhesion to
Human Brain Endothelial Cells. Methods Mol. Biol. 2022,2492, 315–331. [CrossRef]
231.
Varley, J.A.; Strippel, C.; Handel, A.; Irani, S.R. Autoimmune encephalitis: Recent clinical and biological advances. J. Neurol.
2023
,
270, 4118–4131. [CrossRef] [PubMed]
232.
Takeshita, Y.; Fujikawa, S.; Serizawa, K.; Fujisawa, M.; Matsuo, K.; Nemoto, J.; Shimizu, F.; Sano, Y.; Tomizawa-Shinohara, H.;
Miyake, S.; et al. New BBB Model Reveals That IL-6 Blockade Suppressed the BBB Disorder, Preventing Onset of NMOSD. Neurol.
Neuroimmunol. Neuroinflamm. 2021,8, 1076. [CrossRef]
233.
Shimizu, F.; Schaller, K.L.; Owens, G.P.; Cotleur, A.C.; Kellner, D.; Takeshita, Y.; Obermeier, B.; Kryzer, T.J.; Sano, Y.; Kanda, T.;
et al. Glucose-regulated protein 78 autoantibody associates with blood-brain barrier disruption in neuromyelitis optica. Sci.
Transl. Med. 2017,9, aai9111. [CrossRef] [PubMed]
234.
Al-Obaidi, M.M.J.; Desa, M.N.M. Mechanisms of Blood Brain Barrier Disruption by Different Types of Bacteria, and Bacterial-Host
Interactions Facilitate the Bacterial Pathogen Invading the Brain. Cell. Mol. Neurobiol.
2018
,38, 1349–1368. [CrossRef] [PubMed]
235.
Brown, J.A.; Codreanu, S.G.; Shi, M.; Sherrod, S.D.; Markov, D.A.; Neely, M.D.; Britt, C.M.; Hoilett, O.S.; Reiserer, R.S.; Samson,
P.C.; et al. Metabolic consequences of inflammatory disruption of the blood-brain barrier in an organ-on-chip model of the
human neurovascular unit. J. Neuroinflamm. 2016,13, 306. [CrossRef] [PubMed]
236.
Kim, B.J.; Shusta, E.V.; Doran, K.S. Past and Current Perspectives in Modeling Bacteria and Blood-Brain Barrier Interactions.
Front. Microbiol. 2019,10, 1336. [CrossRef]
237.
Devraj, G.; Guérit, S.; Seele, J.; Spitzer, D.; Macas, J.; Khel, M.I.; Heidemann, R.; Braczynski, A.K.; Ballhorn, W.; Günther, S.; et al.
HIF-1
α
is involved in blood-brain barrier dysfunction and paracellular migration of bacteria in pneumococcal meningitis. Acta
Neuropathol. 2020,140, 183–208. [CrossRef]
238.
Caporarello, N.; Olivieri, M.; Cristaldi, M.; Scalia, M.; Toscano, M.A.; Genovese, C.; Addamo, A.; Salmeri, M.; Lupo, G.; Anfuso,
C.D. Blood-Brain Barrier in a Haemophilus influenzae Type a In Vitro Infection: Role of Adenosine Receptors A2A and A2B. Mol.
Neurobiol. 2018,55, 5321–5336. [CrossRef]
239.
Brown, G.C.; Camacho, M.; Williams-Gray, C.H. The Endotoxin Hypothesis of Parkinson’s Disease. Mov. Disord.
2023
. [CrossRef]
240. McArthur, S. Regulation of Physiological Barrier Function by the Commensal Microbiota. Life 2023,13, 396. [CrossRef]
241.
Sasso, J.M.; Ammar, R.M.; Tenchov, R.; Lemmel, S.; Kelber, O.; Grieswelle, M.; Zhou, Q.A. Gut Microbiome-Brain Alliance: A
Landscape View into Mental and Gastrointestinal Health and Disorders. ACS Chem. Neurosci. 2023,14, 1717–1763. [CrossRef]
242.
Shahbazi, A.; Sepehrinezhad, A.; Vahdani, E.; Jamali, R.; Ghasempour, M.; Massoudian, S.; Sahab Negah, S.; Larsen, F.S. Gut
Dysbiosis and Blood-Brain Barrier Alteration in Hepatic Encephalopathy: From Gut to Brain. Biomedicines
2023
,11, 1272.
[CrossRef]
243.
Kim, M.H.; van Noort, D.; Sung, J.H.; Park, S. Organ-on-a-Chip for Studying Gut-Brain Interaction Mediated by Extracellular
Vesicles in the Gut Microenvironment. Int. J. Mol. Sci. 2021,22, 13513. [CrossRef] [PubMed]
244.
Raimondi, I.; Izzo, L.; Tunesi, M.; Comar, M.; Albani, D.; Giordano, C. Organ-On-A-Chip in vitro Models of the Brain and the
Blood-Brain Barrier and Their Value to Study the Microbiota-Gut-Brain Axis in Neurodegeneration. Front. Bioeng. Biotechnol.
2019,7, 435. [CrossRef] [PubMed]
245.
Hall, V.; Bendtsen, K.M.S. Getting closer to modeling the gut-brain axis using induced pluripotent stem cells. Front. Cell Dev. Biol.
2023,11, 1146062. [CrossRef] [PubMed]
246.
Kim, J.; Lee, K.T.; Lee, J.S.; Shin, J.; Cui, B.; Yang, K.; Choi, Y.S.; Choi, N.; Lee, S.H.; Lee, J.H.; et al. Fungal brain infection modelled
in a human-neurovascular-unit-on-a-chip with a functional blood-brain barrier. Nat. Biomed. Eng.
2021
,5, 830–846. [CrossRef]
[PubMed]
247.
Krasemann, S.; Haferkamp, U.; Pfefferle, S.; Woo, M.S.; Heinrich, F.; Schweizer, M.; Appelt-Menzel, A.; Cubukova, A.; Barenberg,
J.; Leu, J.; et al. The blood-brain barrier is dysregulated in COVID-19 and serves as a CNS entry route for SARS-CoV-2. Stem Cell
Rep. 2022,17, 307–320. [CrossRef]
248.
Kase, Y.; Sonn, I.; Goto, M.; Murakami, R.; Sato, T.; Okano, H. The original strain of SARS-CoV-2, the Delta variant, and the
Omicron variant infect microglia efficiently, in contrast to their inability to infect neurons: Analysis using 2D and 3D cultures.
Exp. Neurol. 2023,363, 114379. [CrossRef]
249.
Andrews, M.G.; Mukhtar, T.; Eze, U.C.; Simoneau, C.R.; Ross, J.; Parikshak, N.; Wang, S.; Zhou, L.; Koontz, M.; Velmeshev, D.;
et al. Tropism of SARS-CoV-2 for human cortical astrocytes. Proc. Natl. Acad. Sci. USA 2022,119, e2122236119. [CrossRef]
250.
Rhea, E.M.; Logsdon, A.F.; Hansen, K.M.; Williams, L.M.; Reed, M.J.; Baumann, K.K.; Holden, S.J.; Raber, J.; Banks, W.A.; Erickson,
M.A. The S1 protein of SARS-CoV-2 crosses the blood-brain barrier in mice. Nat. Neurosci. 2021,24, 368–378. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 37 of 40
251.
Buzhdygan, T.P.; DeOre, B.J.; Baldwin-Leclair, A.; Bullock, T.A.; McGary, H.M.; Khan, J.A.; Razmpour, R.; Hale, J.F.; Galie, P.A.;
Potula, R.; et al. The SARS-CoV-2 spike protein alters barrier function in 2D static and 3D microfluidic in-vitro models of the
human blood-brain barrier. Neurobiol. Dis. 2020,146, 105131. [CrossRef]
252.
DeOre, B.J.; Tran, K.A.; Andrews, A.M.; Ramirez, S.H.; Galie, P.A. SARS-CoV-2 Spike Protein Disrupts Blood-Brain Barrier
Integrity via RhoA Activation. J. Neuroimmune Pharmacol. 2021,16, 722–728. [CrossRef] [PubMed]
253.
Zhang, L.; Zhou, L.; Bao, L.; Liu, J.; Zhu, H.; Lv, Q.; Liu, R.; Chen, W.; Tong, W.; Wei, Q.; et al. SARS-CoV-2 crosses the blood-brain
barrier accompanied with basement membrane disruption without tight junctions alteration. Signal Transduct. Target. Ther.
2021
,
6, 337. [CrossRef] [PubMed]
254.
Manosso, L.M.; Arent, C.O.; Borba, L.A.; Ceretta, L.B.; Quevedo, J.; Réus, G.Z. Microbiota-Gut-Brain Communication in the
SARS-CoV-2 Infection. Cells 2021,10, 1993. [CrossRef]
255.
Ju, J.; Su, Y.; Zhou, Y.; Wei, H.; Xu, Q. The SARS-CoV-2 envelope protein disrupts barrier function in an in vitro human blood-brain
barrier model. Front. Cell. Neurosci. 2022,16, 897564. [CrossRef]
256.
Boghdeh, N.A.; Risner, K.H.; Barrera, M.D.; Britt, C.M.; Schaffer, D.K.; Alem, F.; Brown, J.A.; Wikswo, J.P.; Narayanan, A.
Application of a Human Blood Brain Barrier Organ-on-a-Chip Model to Evaluate Small Molecule Effectiveness against Venezuelan
Equine Encephalitis Virus. Viruses 2022,14, 2799. [CrossRef]
257.
Mutso, M.; St John, J.A.; Ling, Z.L.; Burt, F.J.; Poo, Y.S.; Liu, X.; Žusinaite, E.; Grau, G.E.; Hueston, L.; Merits, A.; et al. Basic
insights into Zika virus infection of neuroglial and brain endothelial cells. J. Gen. Virol.
2020
,101, 622–634. [CrossRef] [PubMed]
258.
Alimonti, J.B.; Ribecco-Lutkiewicz, M.; Sodja, C.; Jezierski, A.; Stanimirovic, D.B.; Liu, Q.; Haqqani, A.S.; Conlan, W.; Bani-
Yaghoub, M. Zika virus crosses an in vitro human blood brain barrier model. Fluids Barriers CNS 2018,15, 15. [CrossRef]
259.
Papa, M.P.; Meuren, L.M.; Coelho, S.V.A.; Lucas, C.G.d.O.; Mustafá, Y.M.; Lemos Matassoli, F.; Silveira, P.P.; Frost, P.S.; Pezzuto, P.;
Ribeiro, M.R.; et al. Zika Virus Infects, Activates, and Crosses Brain Microvascular Endothelial Cells, without Barrier Disruption.
Front. Microbiol. 2017,8, 2557. [CrossRef]
260.
Chiu, C.F.; Chu, L.W.; Liao, I.C.; Simanjuntak, Y.; Lin, Y.L.; Juan, C.C.; Ping, Y.H. The Mechanism of the Zika Virus Crossing the
Placental Barrier and the Blood-Brain Barrier. Front. Microbiol. 2020,11, 214. [CrossRef]
261.
Chanthick, C.; Suttitheptumrong, A.; Rawarak, N.; Pattanakitsakul, S.N. Transcytosis Involvement in Transport System and
Endothelial Permeability of Vascular Leakage during Dengue Virus Infection. Viruses 2018,10, 69. [CrossRef] [PubMed]
262.
Wang, K.; Zou, S.; Chen, H.; Higazy, D.; Gao, X.; Zhang, Y.; Cao, S.; Cui, M. Zika virus replication on endothelial cells and
invasion into the central nervous system by inhibiting interferon βtranslation. Virology 2023,582, 23–34. [CrossRef]
263.
Leda, A.R.; Bertrand, L.; Andras, I.E.; El-Hage, N.; Nair, M.; Toborek, M. Selective Disruption of the Blood-Brain Barrier by Zika
Virus. Front. Microbiol. 2019,10, 2158. [CrossRef]
264.
Patel, S.; Leibrand, C.R.; Palasuberniam, P.; Couraud, P.O.; Weksler, B.; Jahr, F.M.; McClay, J.L.; Hauser, K.F.; McRae, M. Effects of
HIV-1 Tat and Methamphetamine on Blood-Brain Barrier Integrity and Function In Vitro. Antimicrob. Agents Chemother.
2017
,61.
[CrossRef]
265.
Liao, K.; Niu, F.; Hu, G.; Guo, M.L.; Sil, S.; Buch, S. HIV Tat-mediated induction of autophagy regulates the disruption of ZO-1 in
brain endothelial cells. Tissue Barriers 2020,8, 1748983. [CrossRef]
266.
Banks, W.A.; Robinson, S.M.; Nath, A. Permeability of the blood-brain barrier to HIV-1 Tat. Exp. Neurol.
2005
,193, 218–227.
[CrossRef] [PubMed]
267.
Marino, J.; Maubert, M.E.; Mele, A.R.; Spector, C.; Wigdahl, B.; Nonnemacher, M.R. Functional impact of HIV-1 Tat on cells of the
CNS and its role in HAND. Cell. Mol. Life Sci. CMLS 2020,77, 5079–5099. [CrossRef]
268.
Swingler, M.; Donadoni, M.; Bellizzi, A.; Cakir, S.; Sariyer, I.K. iPSC-derived three-dimensional brain organoid models and
neurotropic viral infections. J. Neurovirol. 2023,29, 121–134. [CrossRef] [PubMed]
269.
Fattakhov, N.; Torices, S.; Becker, S.; Teglas, T.; Naranjo, O.; Toborek, M. A Triple Primary Cell Culture Model of the Human
Blood-Brain Barrier for Studying Ischemic Stroke In Vitro. J. Vis. Exp. JoVE 2022,6, 188. [CrossRef]
270.
Kadir, R.R.A.; Alwjwaj, M.; Bayraktutan, U. Establishment of an In Vitro Model of Human Blood-Brain Barrier to Study the
Impact of Ischemic Injury. Methods Mol. Biol. 2022,2492, 143–155. [CrossRef] [PubMed]
271.
Lyu, Z.; Park, J.; Kim, K.M.; Jin, H.J.; Wu, H.; Rajadas, J.; Kim, D.H.; Steinberg, G.K.; Lee, W. A neurovascular-unit-on-a-chip
for the evaluation of the restorative potential of stem cell therapies for ischaemic stroke. Nat. Biomed. Eng.
2021
,5, 847–863.
[CrossRef]
272.
Nzou, G.; Wicks, R.T.; VanOstrand, N.R.; Mekky, G.A.; Seale, S.A.; El-Taibany, A.; Wicks, E.E.; Nechtman, C.M.; Marrotte,
E.J.; Makani, V.S.; et al. Multicellular 3D Neurovascular Unit Model for Assessing Hypoxia and Neuroinflammation Induced
Blood-Brain Barrier Dysfunction. Sci. Rep. 2020,10, 9766. [CrossRef]
273.
Wevers, N.R.; Nair, A.L.; Fowke, T.M.; Pontier, M.; Kasi, D.G.; Spijkers, X.M.; Hallard, C.; Rabussier, G.; van Vught, R.; Vulto, P.;
et al. Modeling ischemic stroke in a triculture neurovascular unit on-a-chip. Fluids Barriers CNS 2021,18, 59. [CrossRef]
274.
Candelario-Jalil, E.; Dijkhuizen, R.M.; Magnus, T. Neuroinflammation, Stroke, Blood-Brain Barrier Dysfunction, and Imaging
Modalities. Stroke 2022,53, 1473–1486. [CrossRef]
275.
O’Keeffe, E.; Kelly, E.; Liu, Y.; Giordano, C.; Wallace, E.; Hynes, M.; Tiernan, S.; Meagher, A.; Greene, C.; Hughes, S.; et al.
Dynamic Blood-Brain Barrier Regulation in Mild Traumatic Brain Injury. J. Neurotrauma 2020,37, 347–356. [CrossRef]
276.
Hay, J.R.; Johnson, V.E.; Young, A.M.; Smith, D.H.; Stewart, W. Blood-Brain Barrier Disruption Is an Early Event that May Persist
for Many Years after Traumatic Brain Injury in Humans. J. Neuropathol. Exp. Neurol. 2015,74, 1147–1157. [CrossRef] [PubMed]
Int. J. Mol. Sci. 2023,24, 12699 38 of 40
277.
Zargari, M.; Meyer, L.J.; Riess, M.L.; Li, Z.; Barajas, M.B. P188 Therapy in In Vitro Models of Traumatic Brain Injury. Int. J. Mol.
Sci. 2023,24, 3334. [CrossRef] [PubMed]
278. Omelchenko, A.; Singh, N.K.; Firestein, B.L. Current advances in in vitro models of central nervous system trauma. Curr. Opin.
Biomed. Eng. 2020,14, 34–41. [CrossRef] [PubMed]
279.
Schlotterose, L.; Beldjilali-Labro, M.; Schneider, G.; Vardi, O.; Hattermann, K.; Even, U.; Shohami, E.; Haustein, H.D.; Leichtmann-
Bardoogo, Y.; Maoz, B.M. Traumatic Brain Injury in a Well: A Modular Three-Dimensional Printed Tool for Inducing Traumatic
Brain Injury In vitro. Neurotrauma Rep. 2023,4, 255–266. [CrossRef]
280.
Rosas-Hernandez, H.; Cuevas, E.; Escudero-Lourdes, C.; Lantz, S.M.; Gomez-Crisostomo, N.P.; Sturdivant, N.M.; Balachandran,
K.; Imam, S.Z.; Slikker, W.; Paule, M.G.; et al. Characterization of Biaxial Stretch as an In Vitro Model of Traumatic Brain Injury to
the Blood-Brain Barrier. Mol. Neurobiol. 2018,55, 258–266. [CrossRef]
281.
Rosas-Hernandez, H.; Cuevas, E.; Escudero-Lourdes, C.; Lantz, S.M.; Sturdivant, N.M.; Imam, S.Z.; Sarkar, S.; Slikker, W.; Paule,
M.G.; Balachandran, K.; et al. Characterization of uniaxial high-speed stretch as an in vitro model of mild traumatic brain injury
on the blood-brain barrier. Neurosci. Lett. 2018,672, 123–129. [CrossRef]
282.
Salvador, E.; Burek, M.; Förster, C.Y. An In Vitro Model of Traumatic Brain Injury. Methods Mol. Biol.
2018
,1717, 219–227.
[CrossRef] [PubMed]
283.
Wu, Y.H.; Rosset, S.; Lee, T.R.; Dragunow, M.; Park, T.; Shim, V. In Vitro Models of Traumatic Brain Injury: A Systematic Review.
J. Neurotrauma 2021,38, 2336–2372. [CrossRef]
284.
Zehendner, C.M.; Sebastiani, A.; Hugonnet, A.; Bischoff, F.; Luhmann, H.J.; Thal, S.C. Traumatic brain injury results in rapid
pericyte loss followed by reactive pericytosis in the cerebral cortex. Sci. Rep. 2015,5, 13497. [CrossRef] [PubMed]
285.
Löscher, W.; Friedman, A. Structural, Molecular, and Functional Alterations of the Blood-Brain Barrier during Epileptogenesis
and Epilepsy: A Cause, Consequence, or Both? Int. J. Mol. Sci. 2020,21, 591. [CrossRef]
286.
Reiss, Y.; Bauer, S.; David, B.; Devraj, K.; Fidan, E.; Hattingen, E.; Liebner, S.; Melzer, N.; Meuth, S.G.; Rosenow, F.; et al. The
neurovasculature as a target in temporal lobe epilepsy. Brain Pathol. 2023,33, e13147. [CrossRef]
287.
Morin-Brureau, M.; de Bock, F.; Lerner-Natoli, M. Organotypic brain slices: A model to study the neurovascular unit micro-
environment in epilepsies. Fluids Barriers CNS 2013,10, 11. [CrossRef]
288.
Raimondo, J.V.; Heinemann, U.; de Curtis, M.; Goodkin, H.P.; Dulla, C.G.; Janigro, D.; Ikeda, A.; Lin, C.C.K.; Jiruska, P.;
Galanopoulou, A.S.; et al. Methodological standards for in vitro models of epilepsy and epileptic seizures. A TASK1-WG4 report
of the AES/ILAE Translational Task Force of the ILAE. Epilepsia 2017,58 (Suppl. 4), 40–52. [CrossRef] [PubMed]
289.
Vila Verde, D.; de Curtis, M.; Librizzi, L. Seizure-Induced Acute Glial Activation in the in vitro Isolated Guinea Pig Brain. Front.
Neurol. 2021,12, 607603. [CrossRef]
290.
Campos, G.; Fortuna, A.; Falcão, A.; Alves, G. In vitro and in vivo experimental models employed in the discovery and
development of antiepileptic drugs for pharmacoresistant epilepsy. Epilepsy Res. 2018,146, 63–86. [CrossRef]
291.
Le Duigou, C.; Savary, E.; Morin-Brureau, M.; Gomez-Dominguez, D.; Sobczyk, A.; Chali, F.; Milior, G.; Kraus, L.; Meier, J.C.;
Kullmann, D.M.; et al. Imaging pathological activities of human brain tissue in organotypic culture. J. Neurosci. Methods
2018
,
298, 33–44. [CrossRef]
292.
Jones, R.S.G.; da Silva, A.B.; Whittaker, R.G.; Woodhall, G.L.; Cunningham, M.O. Human brain slices for epilepsy research:
Pitfalls, solutions and future challenges. J. Neurosci. Methods 2016,260, 221–232. [CrossRef] [PubMed]
293.
Yamanaka, G.; Takata, F.; Kataoka, Y.; Kanou, K.; Morichi, S.; Dohgu, S.; Kawashima, H. The Neuroinflammatory Role of Pericytes
in Epilepsy. Biomedicines 2021,9, 759. [CrossRef] [PubMed]
294.
Stebbins, M.J.; Gastfriend, B.D.; Canfield, S.G.; Lee, M.S.; Richards, D.; Faubion, M.G.; Li, W.J.; Daneman, R.; Palecek, S.P.; Shusta,
E.V. Human pluripotent stem cell–derived brain pericyte–like cells induce blood-brain barrier properties. Sci. Adv.
2019
,5,
aau7375. [CrossRef] [PubMed]
295.
Arvanitis, C.D.; Ferraro, G.B.; Jain, R.K. The blood-brain barrier and blood-tumour barrier in brain tumours and metastases. Nat.
Rev. Cancer 2020,20, 26–41. [CrossRef]
296.
Blethen, K.E.; Arsiwala, T.A.; Fladeland, R.A.; Sprowls, S.A.; Panchal, D.M.; Adkins, C.E.; Kielkowski, B.N.; Earp, L.E.; Glass, M.J.;
Pritt, T.A.; et al. Modulation of the blood-tumor barrier to enhance drug delivery and efficacy for brain metastases. Neuro-Oncol.
Adv. 2021,3, v133–v143. [CrossRef]
297.
Logun, M.; Zhao, W.; Mao, L.; Karumbaiah, L. Microfluidics in Malignant Glioma Research and Precision Medicine. Adv. Biosyst.
2018,2, 221. [CrossRef]
298.
Piantino, M.; Figarol, A.; Matsusaki, M. Three-Dimensional in vitro Models of Healthy and Tumor Brain Microvasculature for
Drug and Toxicity Screening. Front. Toxicol. 2021,3, 656254. [CrossRef]
299.
Straehla, J.P.; Hajal, C.; Safford, H.C.; Offeddu, G.S.; Boehnke, N.; Dacoba, T.G.; Wyckoff, J.; Kamm, R.D.; Hammond, P.T. A
predictive microfluidic model of human glioblastoma to assess trafficking of blood-brain barrier-penetrant nanoparticles. Proc.
Natl. Acad. Sci. USA 2022,119, e2118697119. [CrossRef] [PubMed]
300.
Gonzales-Aloy, E.; Ahmed-Cox, A.; Tsoli, M.; Ziegler, D.S.; Kavallaris, M. From cells to organoids: The evolution of blood-brain
barrier technology for modelling drug delivery in brain cancer. Adv. Drug Deliv. Rev. 2023,196, 114777. [CrossRef]
301.
Seo, S.; Choi, C.H.; Yi, K.S.; Kim, S.U.; Lee, K.; Choi, N.; Lee, H.J.; Cha, S.H.; Kim, H.N. An engineered neurovascular unit for
modeling neuroinflammation. Biofabrication 2021,13, abf741. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 39 of 40
302.
Zhang, B.; Li, X.; Tang, K.; Xin, Y.; Hu, G.; Zheng, Y.; Li, K.; Zhang, C.; Tan, Y. Adhesion to the Brain Endothelium Selects Breast
Cancer Cells with Brain Metastasis Potential. Int. J. Mol. Sci. 2023,24, 7087. [CrossRef]
303.
Klein, E.; Hau, A.C.; Oudin, A.; Golebiewska, A.; Niclou, S.P. Glioblastoma Organoids: Pre-Clinical Applications and Challenges
in the Context of Immunotherapy. Front. Oncol. 2020,10, 604121. [CrossRef] [PubMed]
304.
Cui, X.; Ma, C.; Vasudevaraja, V.; Serrano, J.; Tong, J.; Peng, Y.; Delorenzo, M.; Shen, G.; Frenster, J.; Morales, R.T.T.; et al.
Dissecting the immunosuppressive tumor microenvironments in Glioblastoma-on-a-Chip for optimized PD-1 immunotherapy.
eLife 2020,9, 52253. [CrossRef] [PubMed]
305.
Spampinato, S.F.; Takeshita, Y.; Obermeier, B. An In Vitro Model of the Blood-Brain Barrier to Study Alzheimer’s Disease: The
Role of β-Amyloid and Its Influence on PBMC Infiltration. Methods Mol. Biol. 2022,2492, 333–352. [CrossRef] [PubMed]
306.
Mackic, J.B.; Stins, M.; McComb, J.G.; Calero, M.; Ghiso, J.; Kim, K.S.; Yan, S.D.; Stern, D.; Schmidt, A.M.; Frangione, B.; et al.
Human blood-brain barrier receptors for Alzheimer’s amyloid-beta 1-40. Asymmetrical binding, endocytosis, and transcytosis at
the apical side of brain microvascular endothelial cell monolayer. J. Clin. Investig. 1998,102, 734–743. [CrossRef]
307.
Giri, R.; Shen, Y.; Stins, M.; Du Yan, S.; Schmidt, A.M.; Stern, D.; Kim, K.S.; Zlokovic, B.; Kalra, V.K. beta-amyloid-induced
migration of monocytes across human brain endothelial cells involves RAGE and PECAM-1. Am. J. Physiol. Cell Physiol.
2000
,
279, C1772–C1781. [CrossRef]
308.
Merlini, M.; Rafalski, V.A.; Rios Coronado, P.E.; Gill, T.M.; Ellisman, M.; Muthukumar, G.; Subramanian, K.S.; Ryu, J.K.; Syme,
C.A.; Davalos, D.; et al. Fibrinogen Induces Microglia-Mediated Spine Elimination and Cognitive Impairment in an Alzheimer’s
Disease Model. Neuron 2019,101, 1099–1108.e6. [CrossRef]
309.
Brandebura, A.N.; Paumier, A.; Onur, T.S.; Allen, N.J. Astrocyte contribution to dysfunction, risk and progression in neurodegen-
erative disorders. Nat. Rev. Neurosci. 2023,24, 23–39. [CrossRef]
310.
Xu, J.; Farsad, H.L.; Hou, Y.; Barclay, K.; Lopez, B.A.; Yamada, S.; Saliu, I.O.; Shi, Y.; Knight, W.C.; Bateman, R.J.; et al. Human
striatal glia differentially contribute to AD- and PD-specific neurodegeneration. Nat. Aging 2023,3, 346–365. [CrossRef]
311.
Schreiner, T.G.; Creang˘a-Murariu, I.; Tamba, B.I.; Lucanu, N.; Popescu, B.O. In Vitro Modeling of the Blood-Brain Barrier for the
Study of Physiological Conditions and Alzheimer’s Disease. Biomolecules 2022,12, 1136. [CrossRef]
312.
Schiffer, D.; Cordera, S.; Cavalla, P.; Migheli, A. Reactive astrogliosis of the spinal cord in amyotrophic lateral sclerosis. J. Neurol.
Sci. 1996,139 (Suppl. 1), 27–33. [CrossRef] [PubMed]
313.
Mayne, K.; White, J.A.; McMurran, C.E.; Rivera, F.J.; de La Fuente, A.G. Aging and Neurodegenerative Disease: Is the Adaptive
Immune System a Friend or Foe? Front. Aging Neurosci. 2020,12, 572090. [CrossRef] [PubMed]
314.
Beers, D.R.; Zhao, W.; Wang, J.; Zhang, X.; Wen, S.; Neal, D.; Thonhoff, J.R.; Alsuliman, A.S.; Shpall, E.J.; Rezvani, K.; et al. ALS
patients’ regulatory T lymphocytes are dysfunctional, and correlate with disease progression rate and severity. JCI Insight
2017
,
2, e89530. [CrossRef] [PubMed]
315.
Miyazaki, K.; Ohta, Y.; Nagai, M.; Morimoto, N.; Kurata, T.; Takehisa, Y.; Ikeda, Y.; Matsuura, T.; Abe, K. Disruption of
neurovascular unit prior to motor neuron degeneration in amyotrophic lateral sclerosis. J. Neurosci. Res.
2011
,89, 718–728.
[CrossRef]
316.
Winkler, E.A.; Sengillo, J.D.; Sullivan, J.S.; Henkel, J.S.; Appel, S.H.; Zlokovic, B.V. Blood-spinal cord barrier breakdown and
pericyte reductions in amyotrophic lateral sclerosis. Acta Neuropathol. 2013,125, 111–120. [CrossRef]
317.
Gois, A.M.; Mendonça, D.M.F.; Freire, M.A.M.; Santos, J.R. In vitro and in vivo models of amyotrophic lateral sclerosis: An
updated overview. Brain Res. Bull. 2020,159, 32–43. [CrossRef]
318.
Arjmand, B.; Kokabi Hamidpour, S.; Rabbani, Z.; Tayanloo-Beik, A.; Rahim, F.; Aghayan, H.R.; Larijani, B. Organ on a Chip: A
Novel in vitro Biomimetic Strategy in Amyotrophic Lateral Sclerosis (ALS) Modeling. Front. Neurol.
2021
,12, 788462. [CrossRef]
[PubMed]
319.
Fellner, L.; Irschick, R.; Schanda, K.; Reindl, M.; Klimaschewski, L.; Poewe, W.; Wenning, G.K.; Stefanova, N. Toll-like receptor 4
is required for α-synuclein dependent activation of microglia and astroglia. Glia 2013,61, 349–360. [CrossRef]
320.
Gorecki, A.M.; Anyaegbu, C.C.; Anderton, R.S. TLR2 and TLR4 in Parkinson’s disease pathogenesis: The environment takes a
toll on the gut. Transl. Neurodegener. 2021,10, 47. [CrossRef]
321.
Dohgu, S.; Takata, F.; Matsumoto, J.; Kimura, I.; Yamauchi, A.; Kataoka, Y. Monomeric
α
-synuclein induces blood-brain barrier
dysfunction through activated brain pericytes releasing inflammatory mediators in vitro. Microvasc. Res.
2019
,124, 61–66.
[CrossRef]
322.
Al-Bachari, S.; Naish, J.H.; Parker, G.J.M.; Emsley, H.C.A.; Parkes, L.M. Blood-Brain Barrier Leakage Is Increased in Parkinson’s
Disease. Front. Physiol. 2020,11, 593026. [CrossRef]
323.
Kuan, W.L.; Bennett, N.; He, X.; Skepper, J.N.; Martynyuk, N.; Wijeyekoon, R.; Moghe, P.V.; Williams-Gray, C.H.; Barker, R.A.
α
-Synuclein pre-formed fibrils impair tight junction protein expression without affecting cerebral endothelial cell function. Exp.
Neurol. 2016,285, 72–81. [CrossRef] [PubMed]
324.
Pediaditakis, I.; Kodella, K.R.; Manatakis, D.V.; Le, C.Y.; Hinojosa, C.D.; Tien-Street, W.; Manolakos, E.S.; Vekrellis, K.; Hamilton,
G.A.; Ewart, L.; et al. Modeling alpha-synuclein pathology in a human brain-chip to assess blood-brain barrier disruption. Nat.
Commun. 2021,12, 5907. [CrossRef] [PubMed]
325.
Linville, R.M.; Nerenberg, R.F.; Grifno, G.; Arevalo, D.; Guo, Z.; Searson, P.C. Brain microvascular endothelial cell dysfunction in
an isogenic juvenile iPSC model of Huntington’s disease. Fluids Barriers CNS 2022,19, 54. [CrossRef]
Int. J. Mol. Sci. 2023,24, 12699 40 of 40
326.
Le Cann, K.; Foerster, A.; Rösseler, C.; Erickson, A.; Hautvast, P.; Giesselmann, S.; Pensold, D.; Kurth, I.; Rothermel, M.; Mattis,
V.B.; et al. The difficulty to model Huntington’s disease in vitro using striatal medium spiny neurons differentiated from human
induced pluripotent stem cells. Sci. Rep. 2021,11, 6934. [CrossRef] [PubMed]
327.
Box, G.E.P. Sampling and Bayes’ Inference in Scientific Modelling and Robustness. J. R. Stat. Soc. Ser. A Gen.
1980
,143, 383.
[CrossRef]
Disclaimer/Publisher’s Note:
The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
... The brain cells receive copper from the BBB and through the BCB, with choroid plexus cells playing a key role in regulating copper entry into the CSF and endothelial cells of intraparenchymal capillaries regulating copper entry into the brain parenchyma [27]. In recent years, the neurovascular unit (NVU) has been introduced in the literature as a collective term including all cell types participating in the integrity of the BBB: brain microvascular endothelial cells (BMECs), pericytes, astrocytes, neurons, microglia, and oligodendrocytes [28]. The highest rate of copper ions' transport into the brain parenchyma, compared to the CSF, clearly indicates that the BBB is the principal site through which copper enters the CNS [29]. ...
Article
Full-text available
Copper is a transition metal essential for growth and development and indispensable for eukaryotic life. This metal is essential to neuronal function: its deficiency, as well as its overload have been associated with multiple neurodegenerative disorders such as Alzheimer’s disease and Wilson’s disease and psychiatric conditions such as schizophrenia, bipolar disorder, and major depressive disorders. Copper plays a fundamental role in the development and function of the human Central Nervous System (CNS), being a cofactor of multiple enzymes that play a key role in physiology during development. In this context, we thought it would be timely to summarize data on alterations in the metabolism of copper at the CNS level that might influence the development of neuropsychiatric symptoms. We present a non-systematic review with the study selection based on the authors’ judgement to offer the reader a perspective on the most significant elements of neuropsychiatric symptoms in Wilson’s disease. We highlight that Wilson’s disease is characterized by marked heterogeneity in clinical presentation among patients with the same mutation. This should motivate more research efforts to disentangle the role of environmental factors in modulating the expression of genetic predisposition to this disorder.
... The blood-brain barrier (BBB) is essential for maintaining the brain's microenvironment. Recent studies have revealed that presenilin is involved in BBB integrity through interactions with a reduced barrier function, reduced drug efflux pump activity, and dimin-ished glucose metabolism [47,48] Dysfunction of the BBB can lead to increased neuroinflammation and infiltration of peripheral immune cells, exacerbating AD pathology [49]. ...
Article
Full-text available
Presenilin, a transmembrane protein primarily known for its role in Alzheimer’s disease (AD) as part of the γ-secretase complex, has garnered increased attention due to its multifaceted functions in various cellular processes. Recent investigations have unveiled a plethora of functions beyond its amyloidogenic role. This review aims to provide a comprehensive overview of presenilin’s diverse roles in AD and other neurodegenerative disorders. It includes a summary of well-known substrates of presenilin, such as its involvement in amyloid precursor protein (APP) processing and Notch signaling, along with other functions. Additionally, it highlights newly discovered functions, such as trafficking function, regulation of ferritin expression, apolipoprotein E (ApoE) secretion, the interaction of ApoE and presenilin, and the Aβ42-to-Aβ40-converting activity of ACE. This updated perspective underscores the evolving landscape of presenilin research, emphasizing its broader impact beyond established pathways. The incorporation of these novel findings accentuates the dynamic nature of presenilin’s involvement in cellular processes, further advancing our comprehension of its multifaceted roles in neurodegenerative disorders. By synthesizing evidence from a range of studies, this review sheds light on the intricate web of presenilin functions and their implications in health and disease.
... Multiple studies in animal models of epilepsy have shown that activating the HMGB1/TLR4 signaling pathway leads to a marked increase in the frequency of seizures (7,27,37,38). The impairment of the blood-brain barrier could result in the release of serum proteins into the brain, resulting in inflammation and glial activation (39)(40)(41). In cases of SE, the elevated expression of GFAP, a marker of glial cells have been observed. ...
Article
Full-text available
Objective. This study investigated several inflammatory markers’ gene and protein expression in status epilepticus (SE) and their correlation with diazepam resistance. Materials and Methods. Peripheral blood samples were collected from 18 adult patients with SE in Cipto Mangunkusumo Central Hospital, consisting of 12 diazepam-responsive and six diazepam-resistant samples, within 72 hours of the onset of the seizure. We collected baseline demographic and clinical data from each subject. Peripheral blood mononuclear cells (PBMCs) were isolated, cultured, stimulated with lipopolysaccharide (LPS) 1 mg/ml, and harvested for RNA isolation. The RNA was used to determine the expression of Human Mobility Group Box 1 (HMGB1), Inter- leukin-6 (IL-6), IL-10, Toll-like Receptor 4 (TLR4), and Glial fibrillary acidic protein (GFAP). In addition, we performed serum protein assay of HMGB1, IL-6, IL-10, TLR4, and GFAP to compare with gene expression. Results. We found a significant differ- ence between the responsive and resistant groups for serum HMGB1 and IL-6 concentration. The mRNA expression of HMGB1 and IL-6 was significantly higher in LPS-stimulated samples in the responsive but not in the resistant groups. The ratio of IL-6 to IL-10 showed a significant difference between LPS and control in the responsive group. Diazepam response was significantly correlated with seizure duration and serum protein concentration of HMGB1. Conclusion. HMGB1 was highly expressed in the resistant group and strongly correlated with diazepam response, and there was a significant increase in HMGB1 mRNA expres- sion in response to LPS stimulation. These findings suggest that targeting HMGB1 may be a promising therapeutic strategy and that HMGB1 levels could be a valuable biomarker for predicting diazepam resistance in SE.
... Microglia are immune cells in the central nervous system (CNS) that play an important role during viral infection and in inflammatory diseases [19]. In addition, brain microvascular endothelial cells are fundamental elements in the blood-brain barrier, a highly selective and semipermeable barrier that prevents the free passage of solutes or blood into the CNS [20]. In the present study we evaluated the susceptibility of human microglia HMC3 cells and brain microvascular endothelial HBEC-5i cells to MAYV and UNAV infection. ...
Preprint
Full-text available
Mayaro (MAYV) and Una (UNAV) are emerging alphaviruses circulating in the Americas. Earlier reports have revealed that MAYV infects different human cell lines, including synovial and dermal fibroblasts, chondrocytes, osteoblasts, astrocytes and pericytes, as well as neural progenitor cells. In this study we evaluated the susceptibility of immortalized human microglia HMC3 cells and brain microvascular endothelial HBEC-5i cells to MAYV and UNAV infection. Cytopathic effects, cell viability, viral progeny yields, and the presence of E1 and nsP1 proteins in HMC3 and HBEC-5i cells infected with several MAYV or UNAV strains were assessed using an inverted microscope, MTT assay, plaque-forming assays, and immunofluorescence or Western blot, respectively. Finally, the expression of immune response genes was analyzed using qRT-PCR. MAYV and UNAV demonstrated strong cytopathic effects and significantly reduced cell viability in HMC3 cells. Moreover, the HMC3 cells were efficiently infected regardless of the virus strains tested, and E1 and nsP1 viral proteins were detected. In contrast, only MAYV appeared to infect HBEC-5i cells, and minimal effects on cell morphology or viability were observed. Furthermore, the MAYV titer and viral protein levels were substantially lower in the infected HBEC-5i cells when compared to those of the infected microglia cells. Finally, unlike UNAV, MAYV elicited a strong expression of specific interferon-stimulated genes in microglia cells, along with pro-inflammatory cytokines implicated in the immune response. Collectively, these findings demonstrate that MAYV and UNAV are capable of infecting relevant human brain cells.
Article
Full-text available
Prime editing shows potential as a precision genome editing technology, as well as the potential to advance the development of next-generation nanomedicine for addressing neurological disorders. However, turning in prime editors (PEs), which are macromolecular complexes composed of CRISPR/Cas9 nickase fused with a reverse transcriptase and a prime editing guide RNA (pegRNA), to the brain remains a considerable challenge due to physiological obstacles, including the blood–brain barrier (BBB). This review article offers an up-to-date overview and perspective on the latest technologies and strategies for the precision delivery of PEs to the brain and passage through blood barriers. Furthermore, it delves into the scientific significance and possible therapeutic applications of prime editing in conditions related to neurological diseases. It is targeted at clinicians and clinical researchers working on advancing precision nanomedicine for neuropathologies.
Article
Full-text available
Gut microbiota includes a vast collection of microorganisms residing within the gastrointestinal tract. It is broadly recognized that the gut and brain are in constant bidirectional communication, of which gut microbiota and its metabolic production are a major component, and form the so-called gut microbiome-brain axis. Disturbances of microbiota homeostasis caused by imbalance in their functional composition and metabolic activities, known as dysbiosis, cause dysregulation of these pathways and trigger changes in the blood-brain barrier permeability, thereby causing pathological malfunctions, including neurological and functional gastrointestinal disorders. In turn, the brain can affect the structure and function of gut microbiota through the autonomic nervous system by regulating gut motility, intestinal transit and secretion, and gut permeability. Here, we examine data from the CAS Content Collection, the largest collection of published scientific information, and analyze the publication landscape of recent research. We review the advances in knowledge related to the human gut microbiome, its complexity and functionality, its communication with the central nervous system, and the effect of the gut microbiome-brain axis on mental and gut health. We discuss correlations between gut microbiota composition and various diseases, specifically gastrointestinal and mental disorders. We also explore gut microbiota metabolites with regard to their impact on the brain and gut function and associated diseases. Finally, we assess clinical applications of gut-microbiota-related substances and metabolites with their development pipelines. We hope this review can serve as a useful resource in understanding the current knowledge on this emerging field in an effort to further solving of the remaining challenges and fulfilling its potential.
Article
Full-text available
In 2015, we wrote a review in The Journal of Neurology summarizing the field of autoantibody-associated neurological diseases. Now, in 2023, we present an update of the subject which reflects the rapid expansion and refinement of associated clinical phenotypes, further autoantibody discoveries, and a more detailed understanding of immunological and neurobiological pathophysiological pathways which mediate these diseases. Increasing awareness around distinctive aspects of their clinical phenotypes has been a key driver in providing clinicians with a better understanding as to how these diseases are best recognized. In clinical practice, this recognition supports the administration of often effective immunotherapies, making these diseases ‘not to miss’ conditions. In parallel, there is a need to accurately assess patient responses to these drugs, another area of growing interest. Feeding into clinical care are the basic biological underpinnings of the diseases, which offer clear pathways to improved therapies toward enhanced patient outcomes. In this update, we aim to integrate the clinical diagnostic pathway with advances in patient management and biology to provide a cohesive view on how to care for these patients in 2023, and the future.
Article
Full-text available
A common neuropsychiatric complication of advanced liver disease, hepatic encephalopathy (HE), impacts the quality of life and length of hospital stays. There is new evidence that gut microbiota plays a significant role in brain development and cerebral homeostasis. Microbiota metabolites are providing a new avenue of therapeutic options for several neurological-related disorders. For instance, the gut microbiota composition and blood-brain barrier (BBB) integrity are altered in HE in a variety of clinical and experimental studies. Furthermore, probiotics, prebiotics, antibiotics, and fecal microbiota transplantation have been shown to positively affect BBB integrity in disease models that are potentially extendable to HE by targeting gut microbiota. However, the mechanisms that underlie microbiota dysbiosis and its effects on the BBB are still unclear in HE. To this end, the aim of this review was to summarize the clinical and experimental evidence of gut dysbiosis and BBB disruption in HE and a possible mechanism.
Article
Full-text available
Traumatic brain injury (TBI) is a major health problem that affects millions of persons worldwide every year among all age groups, mainly young children, and elderly persons. It is the leading cause of death for children under the age of 16 and is highly correlated with a variety of neuronal disorders, such as epilepsy, and neurodegenerative disease, such as Alzheimer's disease or amyotrophic lateral sclerosis. Over the past few decades, our comprehension of the molecular pathway of TBI has improved, yet despite being a major public health issue, there is currently no U.S. Food and Drug Administration-approved treatment for TBI, and a gap remains between these advances and their application to the clinical treatment of TBI. One of the major hurdles for pushing TBI research forward is the accessibility of TBI models and tools. Most of the TBI models require costume-made, complex, and expensive equipment, which often requires special knowledge to operate. In this study, we present a modular, three-dimensional printed TBI induction device, which induces, by the pulse of a pressure shock, a TBI-like injury on any standard cell-culture tool. Moreover, we demonstrate that our device can be used on multiple systems and cell types and can induce repetitive TBIs, which is very common in clinical TBI. Further, we demonstrate that our platform can recapitulate the hallmarks of TBI, which include cell death, decrease in neuronal functionality, axonal swelling (for neurons), and increase permeability (for endothelium). In addition, in view of the continued discussion on the need, benefits, and ethics of the use of animals in scientific research, this in vitro, high-throughput platform will make TBI research more accessible to other labs that prefer to avoid the use of animals yet are interested in this field. We believe that this will enable us to push the field forward and facilitate/accelerate the availability of novel treatments.
Article
Full-text available
Tumor cells metastasize from a primary lesion to distant organs mainly through hematogenous dissemination, in which tumor cell re-adhesion to the endothelium is essential before extravasating into the target site. We thus hypothesize that tumor cells with the ability to adhere to the endothelium of a specific organ exhibit enhanced metastatic tropism to this target organ. This study tested this hypothesis and developed an in vitro model to mimic the adhesion between tumor cells and brain endothelium under fluid shear stress, which selected a subpopulation of tumor cells with enhanced adhesion strength. The selected cells up-regulated the genes related to brain metastasis and exhibited an enhanced ability to transmigrate through the blood–brain barrier. In the soft microenvironments that mimicked brain tissue, these cells had elevated adhesion and survival ability. Further, tumor cells selected by brain endothelium adhesion expressed higher levels of MUC1, VCAM1, and VLA-4, which were relevant to breast cancer brain metastasis. In summary, this study provides the first piece of evidence to support that the adhesion of circulating tumor cells to the brain endothelium selects the cells with enhanced brain metastasis potential.
Article
Background: A protective and regulatory barrier between the blood and the brain is constituted by the blood-brain barrier (BBB), which comprises microvascular endothelial cells providing homeostatic regulation of the central nervous system (CNS). Inflammation compromises the BBB and contributes to many CNS disorders. Anti-inflammatory effects are exerted by glucocorticoids (GCs) on a variety of cells. These GCs include dexamethasone (Dex), which is used for the treatment of inflammatory diseases and recently for the treatment of COVID-19. Aim: The purpose of this study was to determine whether low or high concentrations of Dex can attenuate the inflammatory response induced by lipopolysaccharide (LPS) in the in vitro BBB model. Method: Brain endothelial cells (bEnd.5) were cultured and exposed to LPS (100ng/ml) and subsequently co-treated with Dex to investigate whether selected concentrations of Dex (0.1,5,10,20µM) can modulate the inflammatory effects of LPS on bEnd.5 cells. Cell viability, cell toxicity, and cell proliferation were investigated, as well as the monitoring of membrane permeability (Trans Endothelial Electrical Resistance-TEER), and Enzyme-Linked Immune Assay (ELISA) kits were used to identify and quantify the presence of inflammatory cytokines (TNF- and IL-1β). Result: Dex, at a lower dosage (0.1µM, but not higher doses), was able to attenuate the inflammatory effects of LPS on bEnd.5 cells. Lower doses of Dex (0.1µM) had no detrimental effects on bEnd.5 cells, while higher Dex doses (5-20µM) decreased bEnd.5 viability, increased bEnd.5 cell toxicity, increased bEnd.5 cell monolayer permeability, and increased proinflammatory cytokine secretion. Conclusion: These results indicate that treatment of brain vascular inflammation with low doses of Dex should be advocated, while higher doses promote vascular inflammation.
Article
Microglial cells are the major immune cells of the central nervous system (CNS), and directly react to neurodegeneration, but other immune cell types are also able to react to pathology and can modify the course of neurodegenerative processes. These mainly include monocytes/macrophages and lymphocytes. While these peripheral immune cells were initially considered to act only after infiltrating the CNS, recent evidence suggests that some of them can also act directly from the periphery. We will review the existing and emerging evidence for a role of peripheral immune cells in neurodegenerative diseases, both with and without CNS infiltration. Our focus will be on amyotrophic lateral sclerosis, but we will also compare to Alzheimer's disease and Parkinson's disease to highlight similarities or differences. Peripheral immune cells are easily accessible, and therefore may be an attractive therapeutic target for neurodegenerative diseases. Thus, understanding how these peripheral immune cells communicate with the CNS deserves deeper investigation.
Article
Cytokines are key messengers by which immune cells communicate, and they drive many physiological processes, including immune and inflammatory responses. Early discoveries demonstrated that cytokines, such as the interleukin family members and TNF-α, regulate synaptic scaling and plasticity. Still, we continue to learn more about how these traditional immune system cytokines affect neuronal structure and function. Different cytokines shape synaptic function on multiple levels ranging from fine-tuning neurotransmission, to regulating synapse number, to impacting global neuronal networks and complex behavior. These recent findings have cultivated an exciting and growing field centered on the importance of immune system cytokines for regulating synapse and neural network structure and function. Here, we highlight the latest findings related to cytokines in the central nervous system and their regulation of synapse structure and function. Moreover, we explore how these mechanisms are becoming increasingly important to consider in diseases-especially those with a large neuroinflammatory component.
Article
The endotoxin hypothesis of Parkinson's disease (PD) is the idea that lipopolysaccharide (LPS) endotoxins contribute to the pathogenesis of this disorder. LPS endotoxins are found in, and released from, the outer membrane of Gram-negative bacteria, for example in the gut. It is proposed that gut dysfunction in early PD leads to elevated LPS levels in the gut wall and blood, which promotes both α-synuclein aggregation in the enteric neurons and a peripheral inflammatory response. Communication to the brain via circulating LPS and cytokines in the blood and/or the gut-brain axis leads to neuroinflammation and spreading of α-synuclein pathology, exacerbating neurodegeneration in brainstem nuclei and loss of dopaminergic neurons in the substantia nigra, and manifesting in the clinical symptoms of PD. The evidence supporting this hypothesis includes: (1) gut dysfunction, permeability, and bacterial changes occur early in PD, (2) serum levels of LPS are increased in a proportion of PD patients, (3) LPS induces α-synuclein expression, aggregation, and neurotoxicity, (4) LPS causes activation of peripheral monocytes leading to inflammatory cytokine production, and (5) blood LPS causes brain inflammation and specific loss of midbrain dopaminergic neurons, mediated by microglia. If the hypothesis is correct, then treatment options might include: (1) changing the gut microbiome, (2) reducing gut permeability, (3) reducing circulating LPS levels, or (4) blocking the response of immune cells and microglia to LPS. However, the hypothesis has a number of limitations and requires further testing, in particular whether reducing LPS levels can reduce PD incidence, progression, or severity. © 2023 The Authors. Movement Disorders published by Wiley Periodicals LLC on behalf of International Parkinson and Movement Disorder Society.
Article
Progress in stem cell research has revolutionized the medical field for more than two decades. More recently, the discovery of induced pluripotent stem cells (iPSCs) has allowed for the development of advanced disease modeling and tissue engineering platforms. iPSCs are generated from adult somatic cells by reprogramming them into an embryonic-like state via the expression of transcription factors required for establishing pluripotency. In the context of the central nervous system (CNS), iPSCs have the potential to differentiate into a wide variety of brain cell types including neurons, astrocytes, microglial cells, endothelial cells, and oligodendrocytes. iPSCs can be used to generate brain organoids by using a constructive approach in three-dimensional (3D) culture in vitro. Recent advances in 3D brain organoid modeling have provided access to a better understanding of cell-to-cell interactions in disease progression, particularly with neurotropic viral infections. Neurotropic viral infections have been difficult to study in two-dimensional culture systems in vitro due to the lack of a multicellular composition of CNS cell networks. In recent years, 3D brain organoids have been preferred for modeling neurotropic viral diseases and have provided invaluable information for better understanding the molecular regulation of viral infection and cellular responses. Here we provide a comprehensive review of the literature on recent advances in iPSC-derived 3D brain organoid culturing and their utilization in modeling major neurotropic viral infections including HIV-1, HSV-1, JCV, ZIKV, CMV, and SARS-CoV2.