ArticlePDF AvailableLiterature Review

Promoting Tissue Regeneration by Modulating the Immune System

Authors:

Abstract and Figures

Statement of significance: Most regenerative strategies have not yet proven to be safe or reasonably efficient in the clinic. In addition to stem cells and growth factors, the immune system plays a crucial role in the tissue healing process. Here, we propose that controlling the immune-mediated mechanisms of tissue repair and regeneration may support existing regenerative strategies or could be an alternative to using stem cells and growth factors. The first part of this review we highlight key immune mechanisms involved in the tissue healing process and marks them as potential target for designing regenerative strategies. In the second part, we discuss various approaches using biomaterials and drug delivery systems that aim at modulating the components of the immune system to promote tissue regeneration.
Content may be subject to copyright.
Review article
Promoting tissue regeneration by modulating the immune system
Ziad Julier
a,1
, Anthony J. Park
a,1
, Priscilla S. Briquez
b
, Mikaël M. Martino
a,
a
European Molecular Biology Laboratory Australia, Australian Regenerative Medicine Institute, Monash University, Victoria 3800, Australia
b
Institute for Molecular Engineering, University of Chicago, Chicago, IL 60637, USA
article info
Article history:
Received 31 October 2016
Received in revised form 3 January 2017
Accepted 20 January 2017
Available online 22 January 2017
Keywords:
Regenerative medicine
Immune system
Biomaterials
Drug delivery systems
Cytokines
Inflammation
Scarring
Fibrosis
Macrophages
T cells
abstract
The immune system plays a central role in tissue repair and regeneration. Indeed, the immune response
to tissue injury is crucial in determining the speed and the outcome of the healing process, including the
extent of scarring and the restoration of organ function. Therefore, controlling immune components via
biomaterials and drug delivery systems is becoming an attractive approach in regenerative medicine,
since therapies based on stem cells and growth factors have not yet proven to be broadly effective in
the clinic. To integrate the immune system into regenerative strategies, one of the first challenges is to
understand the precise functions of the different immune components during the tissue healing process.
While remarkable progress has been made, the immune mechanisms involved are still elusive, and there
is indication for both negative and positive roles depending on the tissue type or organ and life stage. It is
well recognized that the innate immune response comprising danger signals, neutrophils and macro-
phages modulates tissue healing. In addition, it is becoming evident that the adaptive immune response,
in particular T cell subset activities, plays a critical role. In this review, we first present an overview of the
basic immune mechanisms involved in tissue repair and regeneration. Then, we highlight various
approaches based on biomaterials and drug delivery systems that aim at modulating these mechanisms
to limit fibrosis and promote regeneration. We propose that the next generation of regenerative therapies
may evolve from typical biomaterial-, stem cell-, or growth factor-centric approaches to an immune-
centric approach.
Statement of Significance
Most regenerative strategies have not yet proven to be safe or reasonably efficient in the clinic. In addi-
tion to stem cells and growth factors, the immune system plays a crucial role in the tissue healing pro-
cess. Here, we propose that controlling the immune-mediated mechanisms of tissue repair and
regeneration may support existing regenerative strategies or could be an alternative to using stem cells
and growth factors. The first part of this review we highlight key immune mechanisms involved in the
tissue healing process and marks them as potential target for designing regenerative strategies. In the
second part, we discuss various approaches using biomaterials and drug delivery systems that aim at
modulating the components of the immune system to promote tissue regeneration.
Ó2017 Acta Materialia Inc. Published by Elsevier Ltd. This is an open access article underthe CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Contents
1. Introduction . . . ....................................................................................................... 14
2. The main actors of the immune response following tissue injury . . . . . . . . . . . . . . . . . .............................................. 14
2.1. Danger signals . . . . . . . . . .......................................................................................... 15
2.2. Neutrophils and mast cells . . . . . . . . . . . . . . . .......................................................................... 16
2.3. Monocytes and macrophages . . . . . . . . . . . . . ................................................... ....................... 16
2.4. Pericytes. . . . . . . . . . . . . . .......................................................................................... 17
http://dx.doi.org/10.1016/j.actbio.2017.01.056
1742-7061/Ó2017 Acta Materialia Inc. Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Corresponding author at: Australian Regenerative Medicine Institute, 15 Innovation Walk, Building 75, Level 1, Martino lab, Victoria 3800, Australia.
E-mail address: mikael.martino@monash.edu (M.M. Martino).
1
These authors contributed equally.
Acta Biomaterialia 53 (2017) 13–28
Contents lists available at ScienceDirect
Acta Biomaterialia
journal homepage: www.elsevier.com/locate/actabiomat
2.5. Dendritic cells . . . ................................................................................................ 18
2.6. T cells. . . . . . . . . . ................................................................................................ 18
2.7. B-cells. . . . . . . . . . ................................................................................................ 18
3. Promoting tissue regeneration by modulating the immune system . . . . . . . . . . .................................................... 18
3.1. Immune modulation by the physicochemical properties of biomaterials . . . . . . . . . . .......................................... 19
3.2. Immune modulation by decellularized ECM . . . . . . . . . . . . . . ............................................................. 20
3.3. Delivery of inflammatory molecules . . . . ............................................................................. 20
3.3.1. SDF-1. . . . . . . . . . . . . . . . . .................................................................................. 20
3.3.2. PGE2 . . . . . . . . . . . . . . . . . .................................................................................. 20
3.4. Delivery of anti-inflammatory molecules ............................................................................. 21
3.4.1. Pro-resolving mediators . . .................................................................................. 21
3.4.2. Inhibitors of TNF-
a
........................................................................................ 21
3.4.3. Inhibitors of the NF-
j
B pathway. . . . . . . . . . . .................................................................. 21
3.4.4. Anti-inflammatory cytokines . . . . . . . . . . . . . . .................................................................. 22
3.4.5. siRNA. . . . . . . . . . . . . . . . . .................................................................................. 22
3.4.6. miRNA . . . . . . . . . . . . . . . . .................................................................................. 22
3.5. Extracellular vesicles. . . . . . . . . . . . . . . . . ............................................................................. 22
3.6. Codelivery of immune modulators and morphogens . . . . . . . ............................................................. 23
4. Conclusion and future perspectives . . . . . . . . . . . . . . . . ....................................................................... 23
Acknowledgements . . . . . . . . . . .......................................................................................... 24
References . .......................................................................................................... 24
1. Introduction
While remarkable progress has been achieved in understanding
the cellular and molecular mechanisms of tissue repair and regen-
eration, it remains unexplained why mammals have a tendency for
imperfect healing and scarring rather than regeneration. There is
ample evidence in different model organisms indicating that the
immune system is crucial to determine the quality of the repair
response, including the extent of scarring, and the restoration of
organ structure and function. A widespread idea derived from find-
ings in diverse species is that the loss of regenerative capacity is
linked to the evolution of immune competence (Fig. 1). Still, there
are many situations where the immune response to tissue injury
promotes tissue healing. Indeed, the relationship between tissue
healing and the immune response is very complex, since there
are both negative and positive roles, depending on the tissue, organ
and life stage (embryonic, neonatal or adult) [1]. The type of
immune response, its duration and the cells involved can drasti-
cally change the outcome of the tissue healing process from incom-
plete healing and repair (i.e. scarring or fibrosis) to complete
restoration (i.e. regeneration).
In regenerative medicine, strategies based on stem cells and
growth factors have not yet proven broadly effective in the clinic.
Here, we propose that immune-mediated mechanisms of tissue
repair and regeneration may support existing regenerative strate-
gies or could be an alternative to using stem cells and growth fac-
tors. In the first part of this review, we present key immune
mechanisms involved in the tissue healing process, in order to
highlight potential targets. In the second part, we discuss various
approaches using biomaterials and drug delivery systems that
aim at modulating the components of the immune system to pro-
mote tissue repair and regeneration.
2. The main actors of the immune response following tissue
injury
An immune response almost always follows tissue damage and
this response is usually resolved within days to weeks after an
injury. The first phase of the immune response involves compo-
nents of the innate immune system, which provide instant defense
against potential pathogens invading the damaged tissue. How-
ever, even in the absence of pathogens, the immune response ini-
tially triggered by danger signals released from damaged tissues
produces a so-called sterile inflammation [2,3]. In many if not all
tissues, the innate immune response strongly modulates the heal-
ing process. For instance, macrophages and their various pheno-
types play a predominant role in the restoration of tissue
homeostasis by clearing away cellular debris, remodeling the
extracellular matrix (ECM), and synthesizing multiple cytokines
and growth factors. The innate immune response is then followed
by the activation of the adaptive immune system. Although this
was originally thought of as a secondary actor in the tissue healing
process, the adaptive immune response to tissue injury most likely
plays a critical role during tissue repair and regeneration, in partic-
ular the activity of T cells. While a large research effort has focused
on how transplanted mesenchymal stem cells (MSCs) modulate T
cell activities and immune tolerance [4,5], our understanding of
how T cells modulate tissue-resident stem cells and the tissue
healing process is just beginning. In the next sections, we review
the roles and importance of the main actors that shape the
immune response following tissue injury.
Regenerative capacity
Immune competence
Neonatal
Fetal
Adult
Fish
Amphibians
Reptiles
Birds
Mammals
Fig. 1. Apparent inverse relationship between regenerative and immune capacities
during evolution or development. Lower vertebrates such as fishes and amphibians
have the ability to completely regenerate many of their tissues. In mammals,
regenerative capacities depend on the developmental stage (i.e. fetal, neonatal, and
adult). Immune competences have increased during evolution and also increase
with life stage in mammals.
14 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
2.1. Danger signals
Directly after tissue injury, a local inflammation is induced in
response to damage-associated molecular patterns (DAMPs, or
alarmins) and pathogen-associated molecular patterns. Endoge-
nous danger signals are typically released from necrotic or stressed
cells and damaged ECM [2,3]. Well-known DAMPs include heat
shock proteins (HSP), monosodium urate, high-mobility group
box protein 1 (HMGB1), extracellular ATP, and nucleic acids includ-
ing mitochondrial DNA. Inflammatory cytokines such as inter-
leukin (IL)-1
a
and IL-33 can also work as DAMPs and are
released passively from necrotic cells. In addition, fragments from
Fig. 2. The main actors of the immune response following tissue injury. (A) Kinetic of immune cell mobilization after tissue injury. Tissue resident cells including tissue-
resident macrophages and
c
dT cells sense tissue damage and trigger the mobilization of other immune cells. Neutrophils are followed by monocytes/macrophages and T cells.
The relative amount of each cell type recruited is not represented. (B) Overview of the initial inflammatory phase following tissue injury. Chronological events are represented
from left to right. Tissue damage is sensed by tissue-resident macrophages via DAMPs. Neutrophils are the first circulating immune cells recruited to the site of injury,
promoting inflammation and monocyte/macrophage recruitment. The inflammation is initially maintained by pro-inflammatory M(IFN-
c
) macrophages, before being
eventually resolved with the help of M(IL-4) macrophages. (C) Overview of the immune mechanisms that can impair tissue healing or drive to scarring and fibrosis. M(IFN-
c
)
macrophages stimulate effector T cells in a positive-feedback loop. Effector T cells may also inhibit the regenerative capacity of tissue resident stem/progenitor cells via
inflammatory cytokines. M(IL-4)-like macrophages with a pro-fibrotic activity encourage ECM protein deposition and subsequent fibrosis (scarring), preventing full
regeneration of the original tissue. Pericytes increase immune cell mobilization and differentiate into scar forming myofibroblasts via growth factors such as TGF-b1. (D)
Overview of the pro-regenerative immune mechanisms. A critical amount of macrophages displaying an anti-inflammatory/anti-fibrotic phenotype (e.g. M(IL-10)-like)
contribute to regeneration through a crosstalk with Tregs, which in turn help sustain the anti-inflammatory/anti-fibrotic phenotype via secretion of anti-inflammatory
cytokines such as IL-10. Tregs may also enhance the regenerative capacity of endogenous stem/progenitor cells through secretion of growth factors. Th2 cells induce/maintain
anti-fibrotic/anti-inflammatory macrophages. Black arrows indicate a differentiation path or secretion of immune modulators/morphogens. Black dashed arrows indicate a
hypothetical differentiation path. Red arrows indicate induction. Blue arrows indicate inhibition.
Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28 15
ECM components such as hyaluronic acid, collagen, elastin, fibro-
nectin and laminin all stimulate inflammation [6,7].
Toll-like receptors (TLRs) and other types of pattern recognition
receptors recognize danger signals and trigger inflammation via
the activation of the transcription factors NF-
j
B or interferon-
regulatory factors. TLRs activate tissue-resident macrophages and
promote the expression of chemoattractants for neutrophils,
monocytes and macrophages (Fig. 2A, B). They also induce the
expression of pro-inflammatory cytokines such as tumor necrosis
factor-
a
(TNF-
a
), IL-1band IL-6 [8,9]. Interestingly, inflammation
in response to necrotic cells is mostly mediated by IL-1 receptor
(IL-1R), which leads to NF-
j
B activation [10]. IL-33 also acts as a
primary danger signal via the ST2 receptor [11]. However, the
dominant danger signal varies in the context of the injury, includ-
ing the location, magnitude, manner of cell death, and time point
after the injury [3].
TLRs and IL-1R1 have been shown to negatively influence the
repair of several tissues [12–22]. For instance, the harmful effect
of TLR4 signaling is apparent in many organs, as seen by the pro-
tection of TLR4-mutant or deficient mice after hepatic, renal, car-
diac, and cerebral ischemia-reperfusion [12–16]. Similarly, IL-1R1
signaling critically regulates infarct healing [17] and disruption
of IL-1 signaling can improve the quality of wound healing
[18,20]. In addition, it has been shown that IL-1R1/MyD88 signal-
ing negatively regulates bone regeneration in the mouse by impair-
ing the regenerative capacities of mouse MSCs [23]. While TLRs
and IL-1R1 seem to be detrimental for many tissues, studies have
shown that skin wound healing is impaired in mice deficient for
various TLRs [24–26]. For example, TLR4 signaling helps wound
healing through stimulation of transforming growth factor-b
(TGF-b) and CC chemokine ligands (CCL)-5 expression [24].
Another endogenous TLR4 agonist, the extra domain A type III
repeat of fibronectin (FNIII EDA) [27], has been reported to be over-
expressed at sites of injury [28,29], and is known to influence skin
repair [30]. For instance, wound healing in FNIII EDA knockout
mice is abnormal [31].
Overall, it is clear that danger signals significantly influence the
healing process at early stages. They are indeed necessary to
induce inflammation, mainly via NF-
j
B, and they are also involved
in neutrophil, monocyte and macrophage mobilization. Yet, in the
case of ischemia-reperfusion and bone regeneration TLR and IL-R1
signaling seem to be detrimental.
2.2. Neutrophils and mast cells
Neutrophils are usually the first inflammatory cell recruited at a
site of injury, enhancing host defense and wound detection while
removing contaminants [32] (Fig. 2A, B). The recruitment of neu-
trophils requires changes on endothelium surface mediated by his-
tamine, cytokines, and chemokines such as C-X-C motif ligand
(CXCL) 8 that are released by tissue resident cells upon pattern
recognition receptor and TLR activation. This will triggers a recruit-
ment cascade involving the capture of free flowing neutrophils, fol-
lowed by their transmigration from the vasculature to the tissue,
facilitated by an increase permeability of the blood vessels at the
injured site [32]. Neutrophils produce antimicrobial substances
and proteases that help kill and degrade potential pathogens
[33]. In addition, they secrete cytokines and growth factors such
as IL-17 and vascular endothelial growth factor (VEGF)-A, which
recruit and activate more neutrophils and other inflammatory
cells, promote angiogenesis, and stimulate proliferation of cells
such as fibroblasts, epithelial cells and keratinocytes (Fig. 2B)
[32–34].
Neutrophils are also able to deploy neutrophil extracellular
traps (NETs) [35], made of chromatin, proteins and enzymes, able
to catch pathogens and either directly kill them or facilitate their
phagocytosis. Yet, the formation of NETs (or NETosis) needs to be
tightly regulated, since NETosis might impair the healing process.
For example, there are evidences of delayed reepithelization in
the case of diabetes where NETosis is enhanced [36]. This is consis-
tent with the observation that neutrophil depletion might acceler-
ate wound closure in diabetic mice [37].
Importantly, neutrophils exhibit anti-inflammatory capacities.
They facilitate the recruitment of monocytes and macrophages,
which phagocytize dying neutrophils and other cellular debris.
Thus, neutrophils promote their own removal and thereby con-
tribute to the resolution of inflammation (Fig. 2B) [32]. For exam-
ple, following myocardial infarction, neutrophils help controlling,
macrophages, polarization, which is a critical step for proper tissue
repair [38]. Therefore, tightly controlling neutrophil mobilization
and functions could be an interesting strategy to promote tissue
repair and regeneration. For instance, pro-resolving mediators
derived from omega 3 fatty acid have the ability to modulate neu-
trophil mobilization as well as their ingestion by macrophages
[39].
Similarly to neutrophils, mast cells participate in the innate
immune response by secreting an array of effector molecules to
recruit eosinophils and monocytes. A large number of mast cells
seem to be detrimental for tissue regeneration. For example, they
enhance acute inflammation and promote scarring in the central
nervous system [40]. Moreover, they persist at high numbers in
chronic wounds [41]. Nevertheless, controlling mast cells to pro-
mote regeneration rather than repair and scarring should be tem-
pered, since mast cells also produce anti-inflammatory mediators,
suggesting alternative and dynamic functions for these cells during
repair [40].
2.3. Monocytes and macrophages
In addition to their role as scavenger cells that phagocytise cel-
lular debris, invading organisms, neutrophils and other apoptotic
cells, macrophages actively regulate the tissue healing process
[42]. A population of tissue macrophages resides in most tissues,
but a large number of macrophages are recruited after tissue
injury, and these often greatly exceed the population of tissue-
resident macrophages [43]. The recruited and resident populations
proliferate and undergo marked phenotypic and functional
changes, in response to the tissue microenvironment. Importantly,
macrophages are a source of various proteases, cytokines, growth
factors, ECM components and soluble mediators promoting tissue
repair, fibrosis, or regeneration [42,44,45].
Macrophages are differentiated from circulating monocytes
which usually arrive at the damaged site 1–3 days after neu-
trophils (Fig. 2A) [46]. Their accumulation will often peak at 4–
7 days after the injury, although elevated accumulations can be
observed up to 21 days [47]. The two main blood monocyte subsets
in the mouse are the Ly6C
hi
CX3CR1
mid
CCR2
+
(CD62L
+
CD43
low
) and
the Ly6C
low
CX3CR1
hi
CCR2
(CD62L
CD43
hi
) monocytes [48]
(human equivalents are the CD14
+
and the CD14
low
CD16
+
mono-
cytes). There is some evidence to suggest that the primary function
of Ly6C
low
cells is to survey endothelial integrity [49,50]. By con-
trast, Ly6C
hi
monocytes represent ‘‘classical monocytes” that are
recruited to sites of inflammation [48].
The two main chemokines/related receptors involved in the
inflammation-dependent recruitment of monocyte subsets from
blood, bone marrow and spleen are CCL2/CCR2 and CX3CL1/
CX3CR1 (Fig. 2B) [51,52]. For instance, fibroblast, epithelial, and
endothelial cells surrounding the injured tissue produce CCL2, in
response to DAMPs and inflammatory cytokines. Interestingly,
depending on the tissue, one or both monocyte subsets are
recruited. For example, only Ly6C
hi
monocytes are recruited from
the circulation in muscle injury models [53,54]. They first acquire
16 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
an inflammatory function and further mature into Ly6C
low
macro-
phages with repair functions. However, after myocardial infarction,
both monocyte subsets appear to home in the injured tissue at dif-
ferent stages of inflammation via CCR2 and CX3CR1, respectively
[55]. The Ly6C
hi
subset first infiltrates the infarcted heart and exhi-
bits inflammatory functions, while the Ly6C
low
subset is recruited
at a later stage and stimulates repair by expressing high amounts
of VEGF-A and by promoting deposition of collagen.
Driving the recruitment of different monocyte populations,
both CCR2 and CX3CR1 appear to be essential for proper healing
in several tissues. For example, Cx3cr1
/
mice display reduced
levels of
a
-smooth muscle actin and collagen, reduced neovascu-
larization as well as delayed healing in skin wounds [56]. Similarly,
the loss of CX3CR1 leads to delayed skeletal muscle repair [57].
Moreover, deficiency in the CCL2–CCR2 axis appears to impair
muscle and skin repair [58]. For instance, Eming and colleagues
have shown that CCR2 is critical for the recruitment of Ly6C
hi
CCR2
+
monocytes to skin wounds, leading to proangiogenic macrophages
crucial for vascularization [59]. Interestingly, the study showed
that macrophages are the main source of VEGF-A in early tissue
repair.
Pro-inflammatory macrophages – the so-called ‘‘M1” macro-
phages – may become polarized towards a variety of alternatively
activated anti-inflammatory ‘‘M2” macrophages [42]. Although
pro-inflammatory and anti-inflammatory macrophages are the
two most frequently investigated phenotypes in studies of tissue
healing, macrophages exhibiting tissue repair, pro-fibrotic, anti-
fibrotic, pro-resolving, and tissue regeneration characteristics are
also commonly mentioned in the literature [42]. Indeed, the M1
and M2 nomenclature originate from in vitro characterization
where the M1 phenotype is produced by exposure to IFN-
c
and
TNF-
a
, while the M2 phenotype is produced by IL-4 or IL-13 [60].
In this review, we adopted the new classification system proposed
by Murray et al. [61] where nomenclature is linked to the activa-
tion standards i.e., M(IFN-
c
), M(IL-4), M(IL-10), and so forth.
Generally, M(IL-4) macrophages are considered as tissue repair
macrophages, since they express several wound healing factors
such as arginase, ECM components and growth factors such as
VEGF-A, platelet-derived growth factor (PDGF) and insulin-like
growth factor (IGF) [42,56,60] (Fig. 2B). Yet, the mechanisms that
drive macrophages to adopt various tissue repair phenotypes
in vivo are still under intense debate [42,60]. Indeed, macrophage
phenotype associated markers may be expressed simultaneously,
making in vivo characterization even more challenging [63].In
addition to cytokines, microRNAs (miRNA), which control messen-
ger RNA translation and degradation (e.g. messenger RNAs of
cytokines and transcription factors), are most likely critical regula-
tors of macrophage polarization [62,63]. More specifically, miR-9,
miR-127, miR-155, and miR-125b have been shown to promote
M(IFN-
c
) polarization while miR-21, miR-124, miR-223, miR-34a,
let-7c, miR-132, miR-146a, and miR-125a-5p support M(IL-4)
polarization in macrophages by targeting various transcription fac-
tors and adaptor proteins [62,63].
While inflammatory macrophages can exacerbate tissue injury
and impair tissue healing, persistent activation or sustained mobi-
lization of M(IL-4) macrophages has been hypothesized to con-
tribute to the development of pathological fibrosis [42] (Fig. 2C).
For example, the pro-fibrotic function of M(IL-4) macrophages
has been attributed to their production and activation of TGF-b1
in models of pulmonary fibrosis [64]. In addition to producing
pro-fibrotic mediators, M(IL-4) macrophages have been shown to
directly enhance the survival and activation of myofibroblasts,
which are key cells producing ECM in all organs [65]. Pro-fibrotic
M(IL-4) macrophages also produce significant amount of matrix
metalloproteinases (MMPs), and some of which serve as essential
drivers of fibrosis [66].
Macrophages may also be anti-inflammatory/anti-fibrotic and
they are thought to be critical for the resolution of most tissue
injury inflammation responses. IL-10 – an immunoregulatory cyto-
kine produced by a variety of cell types, including T helper 2 cells
(Th2), regulatory T (Treg) cells and macrophages – is known to
function as a critical anti-inflammatory mediator [67]. In addition,
anti-inflammatory macrophages regulate the development and
maintenance of IL-10- and TGF-b1-producing Tregs, which
contribute to the resolution of inflammatory responses in multiple
tissues (Fig. 2D) [68]. Nevertheless, beside the expansion of IL-10-
induced anti-inflammatory macrophages, other mechanisms have
also been shown to trigger anti-inflammatory macrophages [42].
For example, IL-6 and IL-21 have been found to enhance IL-4R
expression on macrophages and contribute to the development
of anti-inflammatory and anti-fibrotic macrophage function fol-
lowing stimulation with IL-4 or IL-13 [69,70].
Interestingly, it has been recently demonstrated that macro-
phages are critical for the regeneration (i.e. the full restoration of
the tissue function) of various tissues [71–73]. For example,
Godwin and colleagues found that macrophages are essential for
limb regeneration in adult salamanders [71]. Moreover, mice can
regenerate cardiac tissue until seven days post-birth and it has
been demonstrated that monocytes and macrophages are required
for the cardiac regeneration process. Remarkably, profiling of car-
diac macrophages from regenerating and non-regenerating hearts
indicated that neonatal macrophages have a unique polarization
that does not fit into M(IFN-
c
) or M(IL-4) phenotypes [73].
Importantly, it remains unclear whether an individual macro-
phage (recruited or tissue-resident) is capable of adopting all the
phenotypes at different time in response to the injured tissue
microenvironment, or if distinct subsets of monocytes and macro-
phages are committed to adopt the various phenotypes [42,60]. For
instance, in several tissues such as the central nervous system and
the liver, macrophages switch from a pro-inflammatory phenotype
to a repair phenotype where IL-4, IL-10 and phagocytosis play crit-
ical roles in the conversion [74–76]. In the context of skin injury,
chemokines (e.g. CX3CL1) drives circulating CX3CR1
hi
monocytes
traffic into the damaged site. The CX3CR1
hi
monocytes become M
(IL-4)-like macrophages and secrete factors such as VEGF-A,
TGF-b[56], IL-13, IL-10 and several chemokines [77].
Overall, monocytes and macrophages can exacerbate inflamma-
tion, promote tissue repair and fibrosis, or drive regeneration.
While the detailed mechanisms regulating macrophage functions
during tissue healing are still unclear, their critical role in the
repair and regeneration processes marks them as a primary target
when designing regenerative strategies.
2.4. Pericytes
Pericytes are ubiquitous mural cells of blood microvessels,
which facilitate the initial extravasation of immune cells from
the blood [78]. Pericytes are also a source of stem/progenitor cells
and they secrete multiple growth factors and cytokines, as well as
other soluble mediators [79] (Fig. 2D). For example, they con-
tribute to skeletal muscle regeneration by driving immune cells
to cross the endothelium [80] and they are most likely a source
of myogenic precursors [80]. Pericytes also contribute to tissue
healing by promoting angiogenesis at the damaged site [79]. For
instance, injection of pericytes into mouse cardiac tissue after
infarction improves the healing process, by reducing scar forma-
tion, fibrosis and cardiomyocyte apoptosis via secretion of angio-
genic factors and miRNA [81]. These examples demonstrate that
pericytes are pro-regenerative cells, but they are also source of
scar-forming myofibroblasts in several organs, including skin, liver,
and in the central nervous system [1,79]. In addition, pericytes
interact with the immune cells involved in the scarring process.
Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28 17
For example, they induce the mobilization of Ly6C
hi
monocytes
that further stimulate scar formation by secreting factors such as
TGF-b1, TNF-
a
, and PDGFs (Fig. 2C). These factors induce pericytes
to change their morphology leading to vascular permeability, pro-
liferation and expression of tissue inhibitors of metalloproteinases
(TIMPs) [82]. Therefore, pericytes have the capacity to support
regeneration, but in acute or chronic inflammation their regenera-
tive function can switch to a fibrotic function. Consequently, one
should design strategies to promote the regenerative capacity of
pericytes (i.e. differentiation into functional tissue cells), while
avoiding promotion of their differentiation into myofibroblasts.
2.5. Dendritic cells
In a manner similar to macrophages, dendritic cells (DCs) will
phagocytise particles and process danger signals at the injury site.
Although their precise role during tissue repair and regeneration
remains not fully understood [83], studies show that they play
an important role in the tissue healing process [26,83–85]. For
example, it has been shown that plasmacytoid DCs sense skin
injury via host-derived nucleic acids (recognized by TLR7 and
TLR9) and promote wound healing through type I interferons
[26]. Burn wound closure is also significantly delayed in DC-
deficient mice [83]. The impaired wound healing seems to be asso-
ciated with significant suppression of early cellular proliferation,
granulation tissue formation, wound levels of TGF-b1 and forma-
tion of blood vessels. In addition, in a myocardial infarction model,
DC-depleted mice show impaired ventricular functions and remod-
eling, with particularly high levels of inflammatory cytokines along
with an unbalanced M(IFN-
c
):M(IL-4) macrophages ratio strongly
tilted towards M(IFN-
c
)[85]. DCs most likely act as an
immunoregulator during tissue healing through control of macro-
phages homeostasis.
2.6. T cells
Growing evidence points towards T cells playing a crucial role
in tissue repair and regeneration. While interesting mechanisms
have been revealed, the exact function of the different T cell types
and subsets and their level of accumulation at injury sites are lar-
gely unknown and seem to vary from tissue to tissue. The majority
a
bT cell fraction appears to have both pro- and anti-regenerative
sub-populations. Meanwhile, the minority tissue resident
c
dT cell
fraction has been widely reported as being pro-regenerative [86–
89].
T cells are capable of secreting a diverse range of cytokines and
growth factors, which have beneficial or inhibitory effects on tissue
healing (Fig. 2C, D). In the context of bone, there is evidence that
both CD4
+
(T helper 1, Th1) and CD8
+
(cytotoxic) T cell subsets
inhibit regeneration [90,91]. For example, fracture healing is accel-
erated in Rag1
/
mice (a mouse model without functional T and B
cells) [92] or when CD8
+
T cells are actively depleted [90].Ona
mechanistic level, it has been demonstrated that T cells inhibits
MSC-driven bone formation in the mouse via IFN-
c
and TNF-
a
[91]. Similar research in humans showed that secretion of IFN-
c
and TNF-
a
by effector memory CD8
+
T cells can result in delayed
osteogenesis and fracture healing [90]. On the other hand, studies
have shown that CD4
+
Tregs are critical for the repair and regener-
ation of several tissues including skin [93], bone [91,94], lungs
[95–97], kidney [98,99], skeletal muscle [100,101], and cardiac
muscle [102]. For example, after damage to mouse skeletal
muscles, Tregs can comprise up to 50% of the T cell population
between day 14 and 30 [100]. The presence of Tregs results in
the production of arginase [103] and anti-inflammatory cytokines
such as IL-10 and TGF-b[94]. These secreted factors provide an
anti-inflammatory microenvironment conducive to repair and
polarization of macrophages [94]. Even as conventional T cells
move away, Treg levels remain elevated. This may be because
Tregs that reside in visceral adipose, muscle and lamina propria
express epithelial growth factor receptor (EGF-R) [104,105]. The
expression of EGF-R allows the growth factor amphiregulin
secreted by mast cells to maintain Tregs at the damaged site
[104]. Once present, Tregs proliferate and upregulate amphiregulin
secretion, which is necessary for regeneration [100].
The
c
dT cells are also important in the tissue healing process.
For example, both humans and mice do not heal skin wounds as
fast or effectively in the absence of
c
dT cells [106]. Functionally,
the pro-repair insulin-like growth factor-1 (IGF-1) is produced by
both mouse [107] and human [108]
c
dT cells. In the context of tis-
sue healing, the dendritic epithelial
c
dT cells (DETCs) are the most
well characterized
c
dsubset [88]. DETCs have an unusual
dendritic-like morphology in the mouse skin, and they respond
within hours to skin tissue damage by secreting chemokines and
TNF-
a
to attract macrophages [88]. Additionally, DETCs accelerate
tissue repair by secreting growth factors and cytokines such as IGF-
1, KGF-1 (FGF-7), KGF-2 (FGF-10), IL-22, and IL-17A [88]. For
instance, it has been shown that
c
dT cells peak between 2 and
7 days after bone injury in the mouse and secrete IL-17A, which
enhance osteoblast functions [87]. Additionally,
c
d-derived IL-22
prompts proliferation and migration of epithelial cells in various
tissues [109]. Overall,
c
dT cells play both a central role in recruiting
innate immune cells as well as directly stimulating tissue growth.
We have made significant headway in understanding the
importance of T cells during tissue repair and regeneration, in
particular Tregs and
c
dT cells. Treg and
c
dT cells secreted growth
factors and cytokines are most likely critical for to orchestrate
tissue healing, particularly in skin and muscle. Nevertheless, the
mechanisms by which the different T cell types and their respec-
tive subsets modulate the immune response to tissue injury are
still very elusive. In addition, T cells probably directly interact with
tissue resident stem or progenitor cell populations, and this could
be a useful niche to exploit for designing new regenerative
strategies.
2.7. B-cells
There is little available evidence on the role of B cells in tissue
healing. Given the origin of B cells within the bone marrow, it
would be expected that there would be cross talk between B cells
and bone tissue [110]. For example, IgM
+
B cells are important in
repair by secreting osteoprotegerin to accelerate bone regeneration
[111]. Interestingly, while CD4
+
T cells help upregulate osteoprote-
gerin via the CD40/CD40L pathway, CD8
+
T cells in contrast inhibit
osteoprotegerin expression [111]. As noted above, mice deficient in
both T and B cells have faster bone healing, suggesting depletion of
the adaptive immune system as a promising strategy to augment
bone regeneration. However, we would argue that there is still
much to be discovered regarding the role of B cells in the repair
and regeneration of various tissues.
3. Promoting tissue regeneration by modulating the immune
system
In the first part of this review, we have seen that the immune
system greatly influences tissue repair and regeneration in both
negative and positive fashion. Therefore, controlling the immune
regulations of tissue healing is becoming an attractive avenue in
regenerative medicine, and the design of regenerative strategies
may progress in parallel with our understanding of the crosstalk
between immune components, stem/progenitor cells and the
tissue healing process. In the next sections, we highlight different
18 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
approaches that attempted controlling the immune system to pro-
mote tissue repair or regeneration. In many cases, these
approaches are based on biomaterials or use biomaterials as deliv-
ery systems for immune modulators (Fig. 3).
3.1. Immune modulation by the physicochemical properties of
biomaterials
Implanted biomaterials can have a significant intrinsic effect on
the immune system and macrophage polarization, either promot-
ing or reducing inflammation depending on their physicochemical
properties. The form that the biomaterial takes (solid, hydrogel or
micro/nanoparticles), the level of crosslinking and the degradabil-
ity, the hydrophobicity, the topography, and the nature of the bio-
material (synthetic vs naturally derived) are important parameters
to consider (Fig. 3)[112,113]. Synthetic biomaterials that have
been used to modulate the immune response following tissue
injury are for example poly lactic-co-glycolic acid (PLGA), poly(lac-
tic acid) (PLA), and polyethylene glycol (PEG). Naturally derived
biomaterials are for instance decellularized tissues such as human
or porcine skin or porcine small intestine submucosa (SIS), or fab-
ricated scaffolds made of natural molecules such as collagen, fibrin,
hyaluronic acid, chitosan, alginate or silk.
Illustrating the importance of the biomaterial crosslinking, scaf-
folds with a high level of crosslinking usually drive a predomi-
nantly inflammatory macrophage response [112,113]. For
example, it has been demonstrated that SIS implantation in rat
preferentially induced anti-inflammatory macrophages while a
carbodiimide crosslinked form of SIS induces predominantly
inflammatory macrophages [114]. Similarly, macrophages seeded
on non-crosslinked porcine dermis or non-crosslinked porcine
SIS, produces lower levels of IL-1, IL-6, and IL-8 compared to
macrophages seeded on chemically cross-linked porcine dermis
[115].
The surface chemistry also appears to influence macrophage
adhesion and their cytokine secretion profile. For example, neu-
trally charged hydrophilic-modified polymers have been shown
to promote less macrophage and less foreign body giant cell forma-
tion compared to hydrophobic and ionic surfaces [116]. Although
there were fewer cells on the hydrophilic/neutral surface, the
macrophages were further activated to produce significantly
greater amounts of cytokine (IL-1, IL-6, IL-8, and IL-10) than
hydrophobic and ionic surfaces [116].
When designing a biomaterial, modulating the surface topogra-
phy is an interesting method to regulate the cellular response via
control of cell shape and elasticity. The modulation of macrophage
function, phenotype and polarization to varying topography has
been a subject of research for several decades [112]. Studies on
the role of topography on macrophage polarization strongly sug-
gest an advantage of stimulating macrophage elongation for pro-
moting anti-inflammatory polarization [117]. This can be
achieved by micro-patterning the surface and to control attach-
ment, or could be achieved by patterning macrophage ligands on
the surface to promote elongation of cells [112].
A number of naturally derived biomaterials such as high
molecular weight hyaluronic acid [118] and chitosan [119], which
have radical oxygen species-scavenging properties, have intrinsic
anti-inflammatory properties. Nevertheless, in the case of most
Fig. 3. Strategies based on biomaterials and drug delivery systems to promote tissue regeneration by controlling the immune system. Biomaterial-based strategies aiming at
improving the healing process through immunomodulation can be achieved either by the biomaterial itself and/or by the delivery of immunomodulators. Strategies most
commonly aim either at the delivery of pro-inflammatory modulators to initiate the healing process or the delivery of anti-inflammatory modulators to promote the
resolution phase via anti-inflammatory/anti-fibrotic macrophages. More complex strategies rely on a sequential delivery of pro-inflammatory and anti-inflammatory
molecules to exert a more comprehensive control over the tissue healing process. Black arrows indicate a differentiation path. Red arrows indicate induction. Blue arrows
indicate inhibition. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28 19
biomaterials, loading or functionalization of the biomaterial with
anti-inflammatory molecules is necessary to modulate the inflam-
matory microenvironment. Naturally derived biomaterials such as
collagen and fibrin are ideal for releasing immune modulators
through enzyme-mediated degradation. On the other hand, syn-
thetic materials may allow for increased control over degradation
and release kinetics of therapeutics, with the caveat that the bio-
material itself and its degradation products should cause a mini-
mal response when implanted [115].
3.2. Immune modulation by decellularized ECM
Excised tissues can be processed to separate cells from the ECM,
leaving only a decellularized ECM scaffold. The structure of these
natural scaffolds influences numerous cellular processes and can
be used to create a pro-regenerative environment [120]. Moreover,
with the ECM proteins being highly conserved across species,
xenografts are usually well tolerated [121], limiting the risk of
undesired inflammation which could interfere with the regulation
of the immune environment. Indeed, among other properties,
decellularized ECM has shown to modulate the wound immune
microenvironment through macrophage polarization [114], with
the ability to direct macrophages towards either an M(IFN-
c
)or
M(IL-4) phenotype. This immune modulation usually depends on
the composition and structure of the scaffold. Although the exact
underlying mechanism is still not fully elucidated [122], a recent
study suggests that the effects could be carried out by matrix-
bound microvesicles (MBVs) embedded in the ECM [123]. The
study by Badylak and colleagues showed that MBVs were biologi-
cally active and were partially responsible for the effect of the scaf-
fold. Indeed, after isolation from urinary bladder matrix, MBVs
were able to stimulate neurite extension on neuroblastoma cells.
A potential mediator of this activity are miRNAs present within
MBVs. Interestingly, although a certain number of miRNA were
conserved across multiple MBVs of different source, a significant
amount was tissue-specific and could partially explain the differ-
ent effects induced by decellularized scaffolds depending on their
tissue of origin.
Interestingly, the ability to control inflammation through
macrophage polarization allows xenograft of acellular ECM to be
more beneficial than an autologous transplantation in some cases.
For example, in a model of tendon reconstruction in mice, the use
of decellularized urinary bladder matrix induced a greater migra-
tion of progenitor cells towards reconstructed tendons compared
to autologous grafts [124]. This improved mobilization of progeni-
tor cells seems to be attributed to an anti-inflammatory M(IL-4)-
like response induced by the decellularized ECM scaffold. Indeed,
it has been extensively shown that transplantation of acellular
scaffolds usually results in an M(IL-4)-like response with less scar-
ring compared to cellular scaffolds [122]. In addition, it has been
recently demonstrated that tissue-derived ECM scaffolds induce a
pro-regenerative immune environment through a robust Th2
immune response, which drive macrophage polarization towards
an M(IL-4) phenotype via IL-4 [125].
Importantly, the type of response induced by a decellularized
ECM scaffold highly depends on the source tissue were the ECM
was harvested. Indeed, a study comparing the macrophage
response after being exposed to ECM derived from different types
of tissue showed a very heterogenous behavior [126]. In this study,
SIS, urinary bladder matrix, brain ECM, esophageal ECM, and colo-
nic ECM all induced an M(IL-4) response while dermal ECM
induced an M(IFN-
c
) phenotype. Interestingly, ECMs derived from
liver, and skeletal muscle did not induce a particular macrophage
phenotype.
Decellularized ECMs also present an interesting option for the
delivery of immunomodulatory molecules. For instance, decellu-
larized bones have been used for the sequential release of two
types of cytokines, the pro-inflammatory IFN-
c
and the anti-
inflammatory IL-4 [127]. This sequential release promoted macro-
phage transition from a M(IFN-
c
) to M(IL-4) phenotype and
enhanced vascularization of the bone scaffolds in a murine subcu-
taneous implantation model.
3.3. Delivery of inflammatory molecules
There is a large emphasis on enhancing tissue repair by down-
regulating unwanted inflammation. However, pro-inflammatory
molecules including danger signals and pro-inflammatory cytoki-
nes are necessary to start the tissue healing program. For instance,
the delivery of heat shock protein 70, an endogenous agonist of
TLR2 and TLR4 [128], accelerates wound healing by up-regulating
macrophage-mediated phagocytosis [129]. Similarly, activation of
TLR9 using CpG has been shown to promote skin repair in primates
[130]. These examples demonstrate that the principle of using pro-
inflammatory molecules to treat tissue damage could work in
some cases (Fig. 3). Indeed, the inflammatory chemokine stromal
cell-derived factor-1 (SDF-1, CXCL12) and prostaglandin E2
(PGE2) have been extensively explored in tissue repair and
regeneration.
3.3.1. SDF-1
SDF-1 is an inflammatory and pro-angiogenic chemokine that
has been shown to be very important in the tissue healing process
[131], in particular by its capacity to mobilize progenitor cells
[132]. For instance, both human and mouse MSCs express CXCR4,
a SDF-1 receptor, allowing the cells to traffic towards SDF-1
[133,134]. A large number of studies have used biomaterials such
as silk-collagen [135], gelatin [136], alginate [137], PEGylated fib-
rin [138], poly(lactic-co-glycolic acid) [139], and thiol functional-
ized sP(EO-stat-PO) [140] to deliver SDF-1 in a controlled
manner, both to increase angiogenesis and recruit CXCR4
+
cells,
including macrophages [141], hematopoietic stem cells [132] and
MSCs [139]. Biomaterials delivering SDF-1 have been used for
many tissue types and the usefulness of this strategy has been
demonstrated in tendons [135], cardiac muscle [138,140], skin
[136] and liver models [132]. Nevertheless, one challenge to using
SDF-1 is its sensitivity to protease, as the cytokine is cleaved by
MMP-2 and serine exopeptidase CD26. This unwanted protein
degradation can be overcome by modifying the MMP-2/CD26
cleavage sites or by codelivering enzymes inhibitors such as saxa-
gliptin [132]. A second concern may be that SDF-1 is implicated in
macrophage-driven hypertrophic scar formation. Indeed, cells such
as mouse lung fibrocytes and pro-fibrotic pericytes have also been
shown to traffic towards SDF-1 in vivo [79,97]. Therefore, appropri-
ate SDF-1 dosing is important when designing therapies, to avoid
induction of fibrosis.
3.3.2. PGE2
Prostaglandin E2 (PGE2) is part of a family of pro-inflammatory
lipid molecules known as prostanoids [142]. PGE2 and its multiple
receptors (EP1, EP2, EP3 and EP4) have been involved in both pro-
and anti-regenerative functions. For example, elevated levels of
PGE2 are found in periodontal disease [143]. Conversely, PGE2
can increase bone formation [142,144] and angiogenesis [145].
Within the immune system, PGE2 can induce proliferation of T
cells and cause their apoptosis [142]. Interestingly, while being
pro-inflammatory, PGE2 has also been shown to inhibit prolifera-
tion and skews the immune response to Th2 [142] by inhibiting
IL-12 [146], IFN-
c
[147] and IL-2 [148] secretion by human
lymphocytes.
While PGE2 can be beneficial for tissue healing, PGE2 therapy
requires multiple doses and has significant side effects, making it
20 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
a poor therapeutic [144,149]. As such limiting PGE2 locally using
biomaterial delivery systems would be better than repeated sys-
temic administration of PGE2. For example, PGE2 in a cholesterol
bearing pullulan nanogel was effective in building bone in mice
[150]. Nanogels could be further improved by replacing PGE2 with
an agonist that only binds one of its four receptors. Two studies
using an BMP-2/EP4 agonist combination in either a PEG nanogel
[149] or polylactic acid gel [151] were successful in inducing bone
repair or mineralization respectively in mice. Thus, using an ago-
nist to a specific PGE2 receptor such as EP4 in combination with
growth factors slowly released via a biomaterial may be an effec-
tive therapy.
3.4. Delivery of anti-inflammatory molecules
Although inflammation at the site of tissue injury is necessary
to kick-start the healing response, its resolution is crucial to
advance the healing process and to restore tissue integrity. The
pro-inflammatory function of macrophages is essential during
the early stages of inflammation, but proper tissue healing requires
macrophages to be polarized towards an anti-inflammatory
phenotype. The pro-resolving activity of macrophages notably
includes the development and maintenance of Tregs. Tregs in turn
contribute to creating an anti-inflammatory environment
beneficial to tissue repair and help sustain the anti-inflammatory
phenotypes of macrophages (Fig. 2D). The mechanisms inducing
the passage from a pro-inflammatory state to a resolution state
naturally exist, but therapeutic strategies aiming at promoting this
transition can further improve the healing process (Fig. 3). For
example, polymer particles fabricated from poly (cyclohexane-
1,4-diylacetone dimethylene ketal) were loaded with an inhibitor
of p38, a mitogen-activated protein kinases important for immune
cell activation, to diminish the post-infarction inflammatory
response in the myocardium [152]. In a myocardial infarction
model, the particles significantly reduced superoxide and TNF-
a
production, and resulted in a reduction of fibrosis as well as
improved cardiac function.
3.4.1. Pro-resolving mediators
Resolvins, protectins, lipoxins and maresins secreted by phago-
cytes, are specialized pro-resolving mediators derived from omega
3 fatty acids [153,154], limiting both the recruitment of neu-
trophils and their ingestion by macrophages [39]. For example,
pro-resolving mediators upregulate the expression of CCR5 (a
receptor for inflammatory chemokines such as CCL3 and CCL5)
by senescent neutrophils and activated T cells. Thus, CCR5
+
apop-
totic leukocytes sequester inflammatory chemokines and act as
terminators of their signaling during the resolution of inflamma-
tion [155]. A resolvin-based strategy has already proved to be effi-
cient at promoting wound healing in a model of obese diabetic
mice through enhanced resolution of peritonitis [156]. Similarly,
administration of protectin on wounds in the same diabetic mouse
model also improved reepithelization and the formation of granu-
lation tissue as well as innervation [157]. Injections of resolvin and
lipoxin have also been shown to be able to control the macrophage
polarization induced after a chitosan scaffold implantation [158].
Indeed, although chitosan usually induces inflammatory
macrophages when the degree of acetylation exceeds 15% [159],
injections of lipoxin or resolvin were able to shift the polarization
balance towards a anti-inflammatory phenotype in a mouse air-
pouch model.
3.4.2. Inhibitors of TNF-
a
The pro-inflammatory activity of M(IFN-
c
) macrophages is lar-
gely mediated by the release of TNF-
a
. While this cytokine has
been shown to positively regulate tissue repair and regeneration
in some situations, its excess can impair the healing process. For
example, pathological levels of TNF-
a
may induce osteoclastogen-
esis (via T cell secretion of RANKL which activates RANK on osteo-
clasts) resulting in more bone reabsorption than osteogenesis.
Thus, strategies aiming at blocking the activity of TNF-
a
have been
proposed to diminish the effect of the pro-inflammatory macro-
phages. Local delivery of common painkillers including aspirin
[160], ibuprofen [161] and pentoxifylline [162] have shown
encouraging results in reducing TNF-
a
. For example, simply deliv-
ering aspirin locally with hydroxyapatite/tricalcium phosphate
ceramic particles could reduce TNF-
a
and prevent apoptosis of
transplanted MSCs, resulting in more bone regeneration [91].
Other strategies include directly targeting TNF-
a
with TNF-
a
anti-
bodies. For example, a delivery system based on chitosan/collagen
scaffold has been developed [163], in which a glucose-sensitive
delivery system was capable of releasing TNF-
a
antibodies upon
increase of glucose level in a diabetic rat model, a condition often
associated with alveolar bone destruction and high level of TNF-
a
.
The system successfully reduced inflammation and promoted alve-
olar bone healing. Other studies have used hyaluronic acid as a
delivery vehicle for anti-TNF-
a
. Hyaluronic acid can bind CD44
on macrophages and thus provide the anti-TNF signal directly to
the cell producing the cytokine. For example, hyaluronic acid plus
a monoclonal antibody for TNF-
a
was effective at inducing early
healing in the rats after a burn [164].
Although many studies have focused on inhibition of TNF-
a
as a
therapy to overcome unwanted inflammation and accelerate tissue
healing, it should be noted that TNF-
a
might be a useful cytokine to
help begin the healing process in some tissues. For example, in a
rat model, pre-stimulation of MSCs with TNF-
a
increased their
engraftment to myocardial infarct [165]. Additionally, TNF-
a
enables mobilization of human and mouse MSCs into damaged tis-
sues [166,167]. After bone fracture in the mouse, TNF-
a
levels peak
at 24 h post-injury, and help recruit pro-regenerative cells such as
MSCs [167]. A second wave of TNF-
a
expression peaks at about
four weeks after injury and is necessary for endochondrial bone
formation [167]. In other tissues, TNF-
a
could also play a positive
role as it helps stimulate production of BMP-2 in the context of car-
diac [165] and skin [168] repair.
3.4.3. Inhibitors of the NF-
j
B pathway
Many DAMPs and inflammatory cytokines such as IL-1 and TNF-
a
induce the NF-
j
B pathway and there is a growing body of evi-
dence that inhibiting NF-kB may be a viable option to accelerate
the healing of some tissues. For example, mice deficient in IL-1
receptor antagonist (IL-1Ra) show delayed wound healing due to
higher neutrophil recruitment and subsequent NF-
j
B activation
in fibroblasts resulting in negative regulation of the pro-repair
TGF-bpathway [169]. In addition, targeting NF-
j
B may aid bone
regeneration. For instance, inhibiting NF-
j
B in mouse osteoblasts
can increase bone density in an induced osteoporosis model
[170]. Moreover, it was shown that inflammatory cytokines such
as TNF-
a
and IL-17 reduce osteogenesis of mouse MSCs. These
cytokines impair the Wnt/b-Catenin signaling in MSCs, which is
critical for osteogenesis [171]. Co-delivering MSCs on apatite-
coated PLGA scaffold with a small inhibitor of IKKb(which is a
subunit in the kinase enzyme complex part of the upstream
NF-
j
B signaling) resulted in much more bone formation in vivo
compared to MSCs delivered without inhibitor. Similarly, it was
shown that IL-1bsignaling through the IL1-R1/MyD88 pathway
inhibits mouse MSC proliferation, migration and differentiation
towards osteoblasts [23]. Indeed, MSC response to growth factor
and Wnt signaling were impaired by IL-1b, due to AKT dephospho-
rylation and b-catenin degradation. Mouse calvarial defect treated
with IL1Ra or a MyD88 inhibitor designed to be covalently incorpo-
rated into fibrin hydrogels and to translocate into cells following
Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28 21
hydrogel remodeling by proteases significantly improved bone
regeneration driven by MSCs [23]. Taken together, these studies
indicate that NF-kB inhibition could have a dual positive effect of
reducing inflammation while increasing regeneration driven by
MSCs.
3.4.4. Anti-inflammatory cytokines
Anti-inflammatory cytokines such as IL-4 and IL-10 are critical
for proper tissue repair and regeneration, since they are involved
in M(IFN-
c
) to M(IL-4) macrophage switching [42]. In particular,
mouse models have convincingly shown that IL-10 is necessary
for scar-free healing [172]. Indeed, fetal mice are able to repair
cutaneous damage scar free via IL-10 dependent mechanisms,
and this ability can be replicated in adult mice that overexpress
IL-10 [172]. The same effect seems to happen in the heart. It has
been shown that fibrosis after infarct in the mouse is considerably
reduced, when IL-10 is delivered through heparin-based coacer-
vate [173]. For regenerative medicine purposes, IL-10 has been
mostly delivered using plasmid DNA and virus vectors [176], but
IL-4 is often delivered as a protein to induce M(IL-4) macrophage
polarization. For example, slow release of IL-4 conjugated to a bone
scaffold via biotin-streptavidin can enhance M(IL-4) macrophage
polarization [127]. In a rat model of in vivo peripheral nerve dam-
age, IL-4 delivered via injectable agarose hydrogel was effective in
increasing the number of M(IL-4) macrophages [174]. Interest-
ingly, IL-4 delivery via this method resulted in much more axons
being regrown after three weeks compared to the controls, sug-
gesting controlled release of IL-4 may help with peripheral nerve
repair via M(IL-4) macrophages. In a different context, it was
shown that IL-4 delivery could potentially reduce bone degrada-
tion after joint replacement [175]. The study showed that poly-
ethylene particles could degrade mouse calvarias in vitro, but this
process was ameliorated via addition of IL-4 [175]. This study sug-
gests that incorporating IL-4 or similar anti-inflammatory cytokine
to implanted materials may prevent unwanted side effects of
implants. Overall, the delivery of IL-4 may enhance tissue regener-
ation in various situations via M(IL-4) macrophage induction.
What is remaining to be explored is the direct effect that IL-4 could
have on T cells.
Another interesting anti-inflammatory cytokine is TGF-b1,
which is necessary for tissue repair at the earliest stages [176],
although the molecule has inflammatory or anti-inflammatory
effects depending on the cell type it signals. For example, TGF-b1
can suppress lymphocyte proliferation as well as activity and can
help to induce immune-suppressive Tregs [177]. Nevertheless,
the cytokine is also highly involved in scar formation [176]. How-
ever, TGF-bhas three isoforms (TGF-b1, 2 and 3) and there is evi-
dence showing that TGF-b3 can be harnessed to accelerate
regeneration and avoid scarring [176]. Indeed, TGF-b3 simply
injected alone on incisional wounds in human patients was able
to slightly, but noticeably reduce post-operative scarring [178].
The design of an optimal delivery system for TGF-b3 may therefore
improve its anti-fibrotic capacity in humans.
3.4.5. siRNA
Small interfering RNA (siRNA)-mediated gene silencing offers
an alternative therapeutic strategy to antibodies and chemical-
based inhibitors. A number of studies have demonstrated the
potential of RNA interference to suppress pro-inflammatory path-
ways and inflammatory cytokines [176,179–182]. The major chal-
lenges for therapeutic use of siRNAs are to develop methods for
delivering siRNAs to the desired cell types in vivo and to escape
from the endosomal compartment [182].
PLGA particles have been used to deliver TNF-
a
siRNA for treat-
ing inflammation associated with rheumatoid arthritis. In a mouse
model of rheumatoid arthritis, the particles resulted in a reduction
of TNF-
a
production and inflammation in the joint. Similarly, PLGA
particles have been used to deliver polyetherimide(PEI)-conjugated
Fc
c
RIII-targeting siRNA to reduce inflammation [179]. The system
proved to be efficient in a rat model of temporomandibular joint
inflammation with reduction of IL-1 and IL-6. In a remarkable study,
a lipid nanoparticle was used to deliver a therapeutic siRNA that
reduced the accumulation of CCR2 pro-inflammatory monocytes
to inflamed tissue [180]. The siRNA targeting CCR2 was adminis-
tered systemically, and shown to reduce the infarct size in a myocar-
dial infarction model, reduce inflammatory cells in atherosclerotic
lesion, improve the survival of pancreatic islet allografts, and reduce
tumor volume. Similarly, nanoparticle-based RNA interference that
effectively silences five key adhesion molecules for arterial leuko-
cyte recruitment has been used to prevent complications after acute
myocardial infarction [181]. Simultaneously encapsulating siRNA
targeting intercellular cell adhesion molecules 1 and 2, vascular cell
adhesion molecule 1, and E- and P-selectins into polymeric
endothelial-avid nanoparticles reduced the recruitment of neu-
trophil and monocyte after myocardial infarction into atheroscle-
rotic lesions and decreased matrix degrading plaque protease
activity. The five-gene combination RNA interference also curtailed
leukocyte recruitment to ischemic myocardium. Overall, these
studies emphasize the potential of siRNA as a therapeutic to control
the immune system and to reduce the detrimental effect of exces-
sive inflammation during the tissue healing process.
3.4.6. miRNA
miRNAs play an important role in immunity [183–185] and tis-
sue healing [186–188]. Their ability to regulate the immune sys-
tem on multiple levels is of particular interest here as they
appear to be involved in the development and functions of
hematopoietic stem cells, as well as innate and adaptive immune
cells. For example, miRNAs can direct macrophages polarization
through targeting of the IRF/STAT pathway, promoting inflamma-
tion and its resolution [62]. Moreover, miRNAs can induce Tregs
[189] and regulate many other aspects of the T cell response
[190] via modulation of TCR signaling [191] and T helper cells plas-
ticity [192]. For instance, miR-181a has the ability to initially help
activate mature T cells through increased TCR signaling sensitivity,
but also to later repress this activation through downregulation of
CD69, a promoter of T cell proliferation [196].
Although therapeutic strategies based on the delivery of miR-
NAs are still scarce, studies have shown that their use can be ben-
eficial to tissue healing [193]. For example, in a rat skeletal muscle
injury model, the combined injection of three different miRNAs
improved muscle regeneration while preventing fibrosis [194].
Nevertheless, direct injections of miRNA present limitations such
as in vivo stability or biodistribution, which could be overcome
by the development of advanced delivery systems [193,195].As
for siRNA, biomaterial-based delivery systems are necessary to
optimize the delivery of miRNA [196]. For instance, hydrogels have
been successfully used for ex vivo delivery of miRNAs to cells in 3D
culture [197] and could provide an alternative to soluble injections
for in vivo delivery at a specific site. Nanoparticle are also a good
option for in vivo delivery of miRNA. For example, delivery of
miR-146a using PEI nanoparticles was able to inhibit renal fibrosis
through suppression of the infiltration of F4/80
+
macrophages
[198]. Currently, options are also being pursued to modulate miR-
NAs signaling in vivo, either by overexpression or inhibition. How-
ever, improvements in delivery methods of the modulators are also
required [193].
3.5. Extracellular vesicles
Extracellular vesicles (EVs), which includes exosomes (from the
endosomal compartment), microvesicles (formed by budding of
22 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
the plasma membrane), and apoptotic bodies (from dying cells),
are phosphor lipid vesicles from 30 nm to 1
l
m in diameter, used
as cargo container by cells to exchange biomolecules such as trans-
membrane receptors and genetic information [199]. More specifi-
cally, their payload can include cytokines, morphogens, MMPs,
antigens, DNA, non-coding RNAs, mRNAs, and miRNAs, with the
latter being particularly explored in term of the potential func-
tional roles of exosomes [199]. EVs are released by most cell types
and have been detected in all bodily fluids. Once released in the
extracellular milieu, EVs are taken up by the target cells within
the local microenvironment or carried to distant sites though bio-
logical fluids [200].
EVs are most likely important regulators of immune cell
activity and could therefore modulate tissue repair and regener-
ation via the immune system (Fig. 3)[201,202]. EVs derived
from immune cells have been principally studied in the context
of the immune response itself, in particular for cancer
immunotherapy [203]. In the context of tissue healing, there
are few studies looking into the potential of immune cell-
derived EVs. Nevertheless, EVs from immune cells most likely
have a role in the crosstalk between immunity and tissue haling
[201,202]. On the other hand, the most studied EVs in tissue
repair and regeneration have been MSC-derived EVs, whose
functions include immunomodulation [204]. For example, MSC-
derived EVs have been explored as a treatment for fibrotic liver
disease [205]. It was demonstrated that the delivery of exosomes
into the liver reduced fibrosis through inhibition of the TGF-b1
signaling pathway, as well as through inhibition of epithelial-
to-mesenchymal transition of hepatocytes. This effect was most
likely mediated by miR-125b [206]. MSC-derived exosomes can
also inhibit macrophage activation by suppressing TLR signaling
[207]. In a mouse model of hypoxic pulmonary hypertension, it
was shown that the intravenous delivery of MSC-derived exo-
somes suppresses hypoxic inflammation by inhibition of pro-
proliferative pathways via miR-17 [208].
In most of the animal in vivo studies, EVs have been delivered by
intravenous or intraperitoneal injection every few days [201,202].
Following injection, studies reported very variable half-life for
vesicles clearance, ranging from 10 min to 12 h [199]. However,
there is currently no evidence that EV administration following tis-
sue injury preferentially homes to the damaged sites [201,202].To
overcome this issue, researchers can use delivery systems to con-
trol the release of EVs in situ. For example, a commercially available
hydrogel (HyStem-HP) has been used to deliver MSC-derived EVs
in a rat critical size bone defect model [209]. In addition, EVs can
be regarded as natural drug delivery vehicles and be loaded with
exogenous therapeutic agents. Since EVs are effective natural sys-
tems for polynucleotide (siRNA, miRNA) and protein delivery,
one can engineer EVs surface to deliver specific content to specific
cell types. For instance, siRNA have been delivered by engineered
EVs to suppress pro-inflammatory genes [210]. Targeting was
achieved by engineering dendritic cell-derived exosomes to dis-
play Lamp2b, an exosomal membrane protein, fused to the
neuron-specific RVG peptide and loaded with exogenous siRNA
by electroporation. The therapeutic potential of exosome-
mediated siRNA delivery was demonstrated in mice, by the signif-
icant mRNA and protein knockdown of a therapeutic target in Alz-
heimer’s disease (BACE1). Similar approaches could be used to
engineer EVs, in order to promote tissue regeneration via immune
modulation.
Overall, EVs are certainly a potential therapeutic for promoting
tissue healing via immune modulation. Furthermore, while a
recent study has demonstrated that EVs are able to stay in decellu-
larized tissue scaffolds [126], new researches should explore novel
methods to deliver EVs and to integrate them into other biomate-
rial scaffolds.
3.6. Codelivery of immune modulators and morphogens
Because the tissue healing process involves numerous immune
and morphogenetic signals operating at the same time or sequen-
tially, the delivery of multiple immune and morphogenetic factors
in a spatiotemporal-controlled manner is most likely required to
induce an effective regenerative microenvironment. When deliver-
ing multiple factors, the first difficulty is to understand which fac-
tors should be delivered at which concentration and at what time.
Then, the challenge is to develop systems able to deliver the differ-
ent factors in a spatiotemporally controlled manner.
One interesting approach is to stimulate an M(IFN-
c
) macro-
phage response, which resolves rapidly to transition to a M(IL-4)
response [112]. In that regard, Alvarez et al. reviewed multiple
strategies to sequentially release different molecules that polarize
the response to M(IFN-
c
), then re-polarize to M(IL-4), resulting in
improved healing [211]. For example, decellularized bones were
used as a scaffold and further engineered to sequentially release
IFN-
c
for inducing a M(IFN-
c
) response and IL-4 to transit to a M
(IL-4) response [127]. The authors were able to confirm the
sequential stimulation of both macrophage phenotypes and
observed increased vascularization in functionalized scaffold com-
pared to empty scaffolds in a murine subcutaneous implantation
model.
Codelivering growth factors and immune modulators is also a
promising approach. For instance, platelet rich plasma (PRP),
which contain high level of growth factors, has been delivered in
l-lactic acid grafted gelatin with macrophage attracting micelles
containing sphingosine-1-phosphate agonist (SEW2871) to
increase bone regeneration in mice [212]. Interestingly,
SEW2871-micelles and PRP enhanced the level of TNF-
a
3 days
after application, and increased the anti-inflammatory cytokines
IL-10 at day 10. Using the same system, SDF-1 codelivered with
SEW2871-micelles was able to more than double the rate of mouse
skin wound closure. Most likely, the combination of SDF-1/
SEW2871 was able to recruit both MSCs and macrophages to the
damaged skin [134]. A similar approach opted for dual delivery
of FGF-2 and IL-10 via poly(ethylene argininylaspartate diglyceride
(PEAD)/heparin coacervate biomaterial in a mouse myocardial
infarct model [173]. The study showed that the combination of
FGF-2 and IL-10 is more effective than delivering the growth factor
and the cytokine alone. Heart function was restored and fibrosis
was significantly reduced. In the context of bone, codelivery of
BMP-2 and EP4 agonist in either a PEG nanogel [149] or polylactic
acid gel [151] was successful in inducing bone regeneration.
4. Conclusion and future perspectives
A remarkable number of immune cells and immune modulators
participate in all phases of the tissue healing process. Therefore
controlling the immune system, in particular key immune cell sub-
sets, is a very plausible strategy to promote tissue regeneration.
However, the complex mechanisms by which the immune system
orchestrates various organs and tissues are still vastly unknown.
For instance, while neutrophils are the first circulating immune
cells mobilized after tissue injury, their role in the healing process
has been somewhat overlooked. They are likely involved in macro-
phage polarization, although we still do not know exactly in what
way. Controlling neutrophil mobilization and functions could be an
interesting strategy to promote tissue regeneration.
Macrophages have shown to be critical during most phases of
the tissue healing process, but the mechanisms by which they
change phenotypes to stimulate tissue repair, fibrosis or full regen-
eration remain unclear. Thus, further effort is required to under-
stand the contributions of the different macrophage populations
Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28 23
and activation states in multiple organ systems. For instance,
inflammatory macrophages mature into anti-inflammatory macro-
phages in certain type of tissues, while a distinct population of
anti-inflammatory macrophages is mobilized in others. Therefore,
depending on the tissue or organ targeted, one could develop
regenerative strategies aimed at stimulating macrophage polariza-
tion or aiming at recruiting pro-wound healing macrophage sub-
sets. The current approaches points towards driving macrophages
polarization to M(IL-4)-like phenotypes, using a variety of immune
modulators delivered through biomaterials and drug delivery
systems. However, one should remember that, although large
numbers of M(IFN-
c
) macrophages exacerbate tissue injury, per-
sistent activation or sustained recruitment of particular M(IL-4)
macrophage subsets contribute to the development of pathological
fibrosis. One of the keys could be to reveal the exact mechanisms
that drive the expansion of anti-inflammatory/anti-fibrotic macro-
phages in vivo and further implement these mechanisms into
regenerative strategies. Perhaps, the answer will come from the
situation in neonates, which display macrophage populations with
pro-regenerative capacities distinct from the adult. For instance,
altering key macrophage transcription factors that stabilize or
induce these particular pro-regenerative populations could hold
promise to stimulate regeneration.
As with macrophage subsets, there is growing evidence that T
cell subsets can have both anti-regenerative and pro-regenerative
properties. Yet, the mechanisms driving T cell mobilization, activa-
tion and conversion at sites of tissue injury are still largely
unknown. We have seen that Tregs are immune-suppressive and
particular Treg populations are pro-regenerative. Th2 and
c
dT cells
could also be critical for inducing a pro-regenerative environment.
Therefore, biomaterials and the delivery of immunomodulators
could be exploited to modulate T cell activities and promote regen-
eration. One could induce T cell conversion into Tregs from con-
ventional T cells recruited at a site of injury or promote the
mobilization of natural Tregs using chemokines. For example,
CCR4, the receptor for CCL22 chemokine, is typically expressed
more on Tregs compared to conventional T cells [213,214]. One
study was able to successfully reduce inflammation and periodon-
tal disease in mice and dogs via injection of CCL22 microparticle
powder [214]. Other options could include increasing the amount
of anti-inflammatory cytokines typically produced by Tregs, Th2
cells and M(IL-4) macrophages such as IL-4, IL-10 or IL-13.
The aging of the immune system is also a parameter that could
be considered when designing regenerative therapies. Indeed, the
baseline macrophage polarization states and the activity of T cell
subsets may be affected by patient characteristics [42,215]. Along
with differences in the number of stem/progenitor cells in neo-
nates and adults, there is now increasing evidence for changes in
macrophage phenotypes and T cell activities with age and diseases.
Therefore, alterations in the crosstalk between tissue-resident or
transplanted stem/progenitor cells and macrophages and/or T cells
could have major impacts on tissue regeneration after injury and in
aging.
Today, ample evidence suggests that an active control of the
immune system is a very plausible therapeutic strategy to induce
tissue regeneration. However, because we still have sparse knowl-
edge about the immune mechanisms modulating the tissue healing
process, one of the main challenges is to target the right immune
cell populations and pathways for the particular tissue or organ
that needs to be regenerated. Then, the challenge is to engineer
efficient biomaterial and delivery system platforms for controlling
the immune-mediated mechanisms of tissue healing. The next
generation of regenerative strategies may evolve from typical
biomaterial-, stem cell-, or growth factor-centric approaches to
an immune-centric approach, seeking to control the immune sys-
tem as a means of promoting regrowth of tissues and organs.
Acknowledgements
This work was supported in part by the research grant of Astel-
las Foundation for Research on Metabolic Disorders to M.M.M.
References
[1] S.J. Forbes, N. Rosenthal, Preparing the ground for tissue regeneration: from
mechanism to therapy, Nat. Med. 20 (2014) 857–869.
[2] S. Hirsiger, H.P. Simmen, C.M. Werner, G.A. Wanner, D. Rittirsch, Danger
signals activating the immune response after trauma, Mediators Inflamm.
2012 (2012) 315941.
[3] H. Kono, A. Onda, T. Yanagida, Molecular determinants of sterile
inflammation, Curr. Opin. Immunol. 26 (2014) 147–156.
[4] Y. Shi, G. Hu, J. Su, W. Li, Q. Chen, P. Shou, C. Xu, X. Chen, Y. Huang, Z. Zhu, X.
Huang, X. Han, N. Xie, G. Ren, Mesenchymal stem cells: a new strategy for
immunosuppression and tissue repair, Cell Res. 20 (2010) 510–518.
[5] Y. Shi, J. Su, A.I. Roberts, P. Shou, A.B. Rabson, G. Ren, How mesenchymal stem
cells interact with tissue immune responses, Trends Immunol. 33 (2012) 136–
143.
[6] Y. Okamura, M. Watari, E.S. Jerud, D.W. Young, S.T. Ishizaka, J. Rose, J.C. Chow,
J.F. Strauss, The extra domain A of fibronectin activates Toll-like receptor 4, J.
Biol. Chem. 276 (2001) 10229–10233.
[7] T.L. Adair-Kirk, R.M. Senior, Fragments of extracellular matrix as mediators of
inflammation, Int. J. Biochem. Cell Biol. 40 (2008) 1101–1110.
[8] T. Kawai, S. Akira, The role of pattern-recognition receptors in innate
immunity: update on Toll-like receptors, Nat. Immunol. 11 (2010) 373–384.
[9] O. Takeuchi, S. Akira, Pattern recognition receptors and inflammation, Cell
140 (2010) 805–820.
[10] C.J. Chen, H. Kono, D. Golenbock, G. Reed, S. Akira, K.L. Rock, Identification of a
key pathway required for the sterile inflammatory response triggered by
dying cells, Nat. Med. 13 (2007) 851–856.
[11] C. Cayrol, J.P. Girard, IL-33: an alarmin cytokine with crucial roles in innate
immunity, inflammation and allergy, Curr. Opin. Immunol. 31 (2014) 31–37.
[12] J. Oyama, C. Blais Jr., X. Liu, M. Pu, L. Kobzik, R.A. Kelly, T. Bourcier, Reduced
myocardial ischemia-reperfusion injury in toll-like receptor 4-deficient mice,
Circulation 109 (2004) 784–789.
[13] A. Tsung, R. Sahai, H. Tanaka, A. Nakao, M.P. Fink, M.T. Lotze, H. Yang, J. Li, K.J.
Tracey, D.A. Geller, T.R. Billiar, The nuclear factor HMGB1 mediates hepatic
injury after murine liver ischemia-reperfusion, J. Exp. Med. 201 (2005) 1135–
1143.
[14] H. Wu, G. Chen, K.R. Wyburn, J. Yin, P. Bertolino, J.M. Eris, S.I. Alexander, A.F.
Sharland, S.J. Chadban, TLR4 activation mediates kidney ischemia/reperfusion
injury, J. Clin. Invest. 117 (2007) 2847–2859.
[15] S.C. Tang, T.V. Arumugam, X. Xu, A. Cheng, M.R. Mughal, D.G. Jo, J.D. Lathia, D.
A. Siler, S. Chigurupati, X. Ouyang, T. Magnus, S. Camandola, M.P. Mattson,
Pivotal role for neuronal Toll-like receptors in ischemic brain injury and
functional deficits, Proc. Natl. Acad. Sci. U.S.A. 104 (2007) 13798–13803.
[16] K. Hyakkoku, J. Hamanaka, K. Tsuruma, M. Shimazawa, H. Tanaka, S. Uematsu,
S. Akira, N. Inagaki, H. Nagai, H. Hara, Toll-like receptor 4 (TLR4), but not TLR3
or TLR9, knock-out mice have neuroprotective effects against focal cerebral
ischemia, Neuroscience 171 (2010) 258–267.
[17] M. Bujak, M. Dobaczewski, K. Chatila, L.H. Mendoza, N. Li, A. Reddy, N.G.
Frangogiannis, Interleukin-1 receptor type I signaling critically regulates
infarct healing and cardiac remodeling, Am. J. Pathol. 173 (2008) 57–67.
[18] A.A. Thomay, J.M. Daley, E. Sabo, P.J. Worth, L.J. Shelton, M.W. Harty, J.S.
Reichner, J.E. Albina, Disruption of interleukin-1 signaling improves the
quality of wound healing, Am. J. Pathol. 174 (2009) 2129–2136.
[19] T. Shichita, E. Hasegawa, A. Kimura, R. Morita, R. Sakaguchi, I. Takada, T.
Sekiya, H. Ooboshi, T. Kitazono, T. Yanagawa, T. Ishii, H. Takahashi, S. Mori, M.
Nishibori, K. Kuroda, S. Akira, K. Miyake, A. Yoshimura, Peroxiredoxin family
proteins are key initiators of post-ischemic inflammation in the brain, Nat.
Med. 18 (2012) 911–917.
[20] R.E. Mirza, M.M. Fang, W.J. Ennis, T.J. Koh, Blocking interleukin-1beta induces
a healing-associated wound macrophage phenotype and improves healing in
type 2 diabetes, Diabetes 62 (2013) 2579–2587.
[21] S. Bhattacharyya, Z. Tamaki, W. Wang, M. Hinchcliff, P. Hoover, S. Getsios, E.S.
White, J. Varga, FibronectinEDA promotes chronic cutaneous fibrosis through
Toll-like receptor signaling, Sci. Transl. Med. 6 (2014). 232ra50.
[22] N. Takemura, T. Kawasaki, J. Kunisawa, S. Sato, A. Lamichhane, K. Kobiyama, T.
Aoshi, J. Ito, K. Mizuguchi, T. Karuppuchamy, K. Matsunaga, S. Miyatake, N.
Mori, T. Tsujimura, T. Satoh, Y. Kumagai, T. Kawai, D.M. Standley, K.J. Ishii, H.
Kiyono, S. Akira, S. Uematsu, Blockade of TLR3 protects mice from lethal
radiation-induced gastrointestinal syndrome, Nat. Commun. 5 (2014) 3492.
[23] M.M. Martino, K. Maruyama, G.A. Kuhn, T. Satoh, O. Takeuchi, R. Muller, S.
Akira, Inhibition of IL-1R1/MyD88 signalling promotes mesenchymal stem
cell-driven tissue regeneration, Nat. Commun. 7 (2016) 11051.
[24] H. Suga, M. Sugaya, H. Fujita, Y. Asano, Y. Tada, T. Kadono, S. Sato, TLR4, rather
than TLR2, regulates wound healing through TGF-band CCL5 expression, J.
Dermatol. Sci. 73 (2014) 117–124.
[25] L. Chen, S. Guo, M.J. Ranzer, L.A. DiPietro, Toll-like receptor 4 plays an
essential role in early skin wound healing, J. Invest. Dermatol. (2012).
[26] J. Gregorio, S. Meller, C. Conrad, A. Di Nardo, B. Homey, A. Lauerma, N. Arai, R.
L. Gallo, J. Digiovanni, M. Gilliet, Plasmacytoid dendritic cells sense skin injury
24 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
and promote wound healing through type I interferons, J. Exp. Med. 207
(2010) 2921–2930.
[27] Z. Julier, M.M. Martino, A. De Titta, L. Jeanbart, J.A. Hubbell, The TLR4 agonist
fibronectin extra domain a is cryptic, exposed by ELastase-2; use in a fibrin
matrix cancer vaccine, Sci. Rep. 5 (2015).
[28] W.R. Jarnagin, D.C. Rockey, V.E. Koteliansky, S.-S. Wang, D.M. Bissell,
Expression of variant fibronectins in wound healing: cellular source and
biological activity of the EIIIA segment in rat hepatic fibrogenesis, J. Cell Biol.
127 (1994) 2037–2048.
[29] Y.-F. Liao, P.J. Gotwals, V.E. Koteliansky, D. Sheppard, L. Van De Water, The
EIIIA segment of fibronectin is a ligand for integrins
a
9b1 and
a
4b1 providing
a novel mechanism for regulating cell adhesion by alternative splicing, J. Biol.
Chem. 277 (2002) 14467–14474.
[30] W.S. To, K.S. Midwood, Plasma and cellular fibronectin: distinct and
independent functions during tissue repair, Fibrogenesis Tissue Repair 4
(2011) 21.
[31] A.F. Muro, A.K. Chauhan, S. Gajovic, A. Iaconcig, F. Porro, G. Stanta, F.E. Baralle,
Regulated splicing of the fibronectin EDA exon is essential for proper skin
wound healing and normal lifespan, J. Cell Biol. 162 (2003) 149–160.
[32] E. Kolaczkowska, P. Kubes, Neutrophil recruitment and function in health and
inflammation, Nat. Rev. Immunol. 13 (2013) 159–175.
[33] T.A. Wilgus, S. Roy, J.C. McDaniel, Neutrophils and wound repair: positive
actions and negative reactions, Adv Wound Care 2 (2013) 379–388.
[34] R.L. Zemans, N. Briones, M. Campbell, J. McClendon, S.K. Young, T. Suzuki, I.V.
Yang, S. De Langhe, S.D. Reynolds, R.J. Mason, M. Kahn, P.M. Henson, S.P.
Colgan, G.P. Downey, Neutrophil transmigration triggers repair of the lung
epithelium via beta-catenin signaling, Proc. Natl. Acad. Sci. U.S.A. 108 (2011)
15990–15995.
[35] V. Brinkmann, U. Reichard, C. Goosmann, B. Fauler, Y. Uhlemann, D.S. Weiss,
Y. Weinrauch, A. Zychlinsky, Neutrophil extracellular traps kill bacteria,
Science 303 (2004) 1532–1535.
[36] S.L. Wong, M. Demers, K. Martinod, M. Gallant, Y. Wang, A.B. Goldfine, C.R.
Kahn, D.D. Wagner, Diabetes primes neutrophils to undergo NETosis, which
impairs wound healing, Nat. Med. 21 (2015) 815–819.
[37] J.V. Dovi, L.-K. He, L.A. DiPietro, Accelerated wound closure in neutrophil-
depleted mice, J. Leukoc. Biol. 73 (2003) 448–455.
[38] M. Horckmans, L. Ring, J. Duchene, D. Santovito, M.J. Schloss, M. Drechsler, C.
Weber, O. Soehnlein, S. Steffens, Neutrophils orchestrate post-myocardial
infarction healing by polarizing macrophages towards a reparative
phenotype, Eur. Heart J. (2016). ehw002.
[39] J.M. Schwab, N. Chiang, M. Arita, C.N. Serhan, Resolvin E1 and protectin D1
activate inflammation-resolution programmes, Nature 447 (2007) 869–874.
[40] S.D. Skaper, L. Facci, P. Giusti, Mast cells, glia and neuroinflammation:
partners in crime?, Immunology 141 (2014) 314–327
[41] B.C. Wulff, T.A. Wilgus, Mast cell activity in the healing wound: more than
meets the eye?, Exp Dermatol. 22 (2013) 507–510.
[42] T.A. Wynn, K.M. Vannella, Macrophages in Tissue Repair, Regeneration, and
Fibrosis, Immunity 44 (2016) 450–462.
[43] L.C. Davies, S.J. Jenkins, J.E. Allen, P.R. Taylor, Tissue-resident macrophages,
Nat. Immunol. 14 (2013) 986–995.
[44] Y. Okabe, R. Medzhitov, Tissue biology perspective on macrophages, Nat.
Immunol. 17 (2016) 9–17.
[45] C. Varol, A. Mildner, S. Jung, Macrophages: development and tissue
specialization, Annu. Rev. Immunol. 33 (2015) 643–675.
[46] G.C. Gurtner, S. Werner, Y. Barrandon, M.T. Longaker, Wound Repair Regener.,
Nature 453 (2008) 314–321.
[47] M.L. Novak,E.M. Weinheimer-Haus, T.J. Koh, Macrophage activation and skeletal
muscle healing following traumatic injury, J. Pathol. 232 (2014) 344–355.
[48] F. Ginhoux, S. Jung, Monocytes and macrophages: developmental pathways
and tissue homeostasis, Nat. Rev. Immunol. 14 (2014) 392–404.
[49] C. Auffray, D. Fogg, M. Garfa, G. Elain, O. Join-Lambert, S. Kayal, S. Sarnacki, A.
Cumano, G. Lauvau, F. Geissmann, Monitoring of blood vessels and tissues by
a population of monocytes with patrolling behavior, Science 317 (2007) 666–
670.
[50] L.M. Carlin, E.G. Stamatiades, C. Auffray, R.N. Hanna, L. Glover, G. Vizcay-
Barrena, C.C. Hedrick, H.T. Cook, S. Diebold, F. Geissmann, Nr4a1-dependent
Ly6C(low) monocytes monitor endothelial cells and orchestrate their
disposal, Cell 153 (2013) 362–375.
[51] F. Geissmann, S. Jung, D.R. Littman, Blood monocytes consist of two principal
subsets with distinct migratory properties, Immunity 19 (2003) 71–82.
[52] P. Italiani, D. Boraschi, From monocytes to M1/M2 macrophages:
phenotypical vs. functional differentiation, Front. Immunol. 5 (2014) 514.
[53] L. Arnold, A. Henry, F. Poron, Y. Baba-Amer, N. van Rooijen, A. Plonquet, R.K.
Gherardi, B. Chazaud, Inflammatory monocytes recruited after skeletal
muscle injury switch into antiinflammatory macrophages to support
myogenesis, J. Exp. Med. 204 (2007) 1057–1069.
[54] M.J. Crane, J.M. Daley, O. van Houtte, S.K. Brancato, W.L. Henry Jr., J.E. Albina,
The monocyte to macrophage transition in the murine sterile wound, PLoS
ONE 9 (2014) e86660.
[55] M. Nahrendorf, F.K. Swirski, E. Aikawa, L. Stangenberg, T. Wurdinger, J.L.
Figueiredo, P. Libby, R. Weissleder, M.J. Pittet, The healing myocardium
sequentially mobilizes two monocyte subsets with divergent and
complementary functions, J. Exp. Med. 204 (2007) 3037–3047.
[56] Y. Ishida, J.-L. Gao, P.M. Murphy, Chemokine receptor CX3CR1 mediates skin
wound healing by promoting macrophage and fibroblast accumulation and
function, J. Immunol. 180 (2008) 569–579.
[57] W. Zhao, H. Lu, X. Wang, R.M. Ransohoff, L. Zhou, CX3CR1 deficiency delays
acute skeletal muscle injury repair by impairing macrophage functions,
FASEB J. 30 (2016) 380–393.
[58] L. Arnold, H. Perrin, C.B. de Chanville, M. Saclier, P. Hermand, L. Poupel, E.
Guyon, F. Licata, W. Carpentier, J. Vilar, CX3CR1 deficiency promotes muscle
repair and regeneration by enhancing macrophage ApoE production, Nat.
Commun. 6 (2015).
[59] S. Willenborg, T. Lucas, G. van Loo, J.A. Knipper, T. Krieg, I. Haase, B.
Brachvogel, M. Hammerschmidt, A. Nagy, N. Ferrara, CCR2 recruits an
inflammatory macrophage subpopulation critical for angiogenesis in tissue
repair, Blood 120 (2012) 613–625.
[60] M.L. Novak, T.J. Koh, Macrophage phenotypes during tissue repair, J. Leukoc.
Biol. 93 (2013) 875–881.
[61] P.J. Murray, J.E. Allen, S.K. Biswas, E.A. Fisher, D.W. Gilroy, S. Goerdt, S.
Gordon, J.A. Hamilton, L.B. Ivashkiv, T. Lawrence, M. Locati, A. Mantovani, F.O.
Martinez, J.L. Mege, D.M. Mosser, G. Natoli, J.P. Saeij, J.L. Schultze, K.A. Shirey,
A. Sica, J. Suttles, I. Udalova, J.A. van Ginderachter, S.N. Vogel, T.A. Wynn,
Macrophage activation and polarization: nomenclature and experimental
guidelines, Immunity 41 (2014) 14–20.
[62] K. Essandoh, Y. Li, J. Huo, G.C. Fan, MiRNA-mediated macrophage polarization
and its potential role in the regulation of inflammatory response, Shock 46
(2016) 122–131.
[63] X.Q. Wu, Y. Dai, Y. Yang, C. Huang, X.M. Meng, B.M. Wu, J. Li, Emerging role of
microRNAs in regulating macrophage activation and polarization in immune
response and inflammation, Immunology 148 (2016) 237–248.
[64] L.A. Murray, Q. Chen, M.S. Kramer, D.P. Hesson, R.L. Argentieri, X. Peng, M.
Gulati, R.J. Homer, T. Russell, N. van Rooijen, J.A. Elias, C.M. Hogaboam, E.L.
Herzog, TGF-beta driven lung fibrosis is macrophage dependent and blocked
by Serum amyloid P, Int. J. Biochem. Cell Biol. 43 (2011) 154–162.
[65] J.P. Pradere, J. Kluwe, S. De Minicis, J.J. Jiao, G.Y. Gwak, D.H. Dapito, M.K. Jang,
N.D. Guenther, I. Mederacke, R. Friedman, A.C. Dragomir, C. Aloman, R.F.
Schwabe, Hepatic macrophages but not dendritic cells contribute to liver
fibrosis by promoting the survival of activated hepatic stellate cells in mice,
Hepatology 58 (2013) 1461–1473.
[66] S.K. Madala, J.T. Pesce, T.R. Ramalingam, M.S. Wilson, S. Minnicozzi, A.W.
Cheever, R.W. Thompson, M.M. Mentink-Kane, T.A. Wynn, Matrix
metalloproteinase 12-deficiency augments extracellular matrix degrading
metalloproteinases and attenuates IL-13-dependent fibrosis, J. Immunol. 184
(2010) 3955–3963.
[67] M. Saraiva, A. O’Garra, The regulation of IL-10 production by immune cells,
Nat. Rev. Immunol. 10 (2010) 170–181.
[68] P. Soroosh, T.A. Doherty, W. Duan, A.K. Mehta, H. Choi, Y.F. Adams, Z.
Mikulski, N. Khorram, P. Rosenthal, D.H. Broide, M. Croft, Lung-resident tissue
macrophages generate Foxp3+ regulatory T cells and promote airway
tolerance, J. Exp. Med. 210 (2013) 775–788.
[69] J. Mauer, B. Chaurasia, J. Goldau, M.C. Vogt, J. Ruud, K.D. Nguyen, S. Theurich,
A.C. Hausen, J. Schmitz, H.S. Bronneke, E. Estevez, T.L. Allen, A. Mesaros, L.
Partridge, M.A. Febbraio, A. Chawla, F.T. Wunderlich, J.C. Bruning, Signaling by
IL-6 promotes alternative activation of macrophages to limit endotoxemia
and obesity-associated resistance to insulin, Nat. Immunol. 15 (2014) 423–
430.
[70] J. Pesce, M. Kaviratne, T.R. Ramalingam, R.W. Thompson, J.F. Urban Jr., A.W.
Cheever, D.A. Young, M. Collins, M.J. Grusby, T.A. Wynn, The IL-21 receptor
augments Th2 effector function and alternative macrophage activation, J.
Clin. Invest. 116 (2006) 2044–2055.
[71] J.W. Godwin, A.R. Pinto, N.A. Rosenthal, Macrophages are required for adult
salamander limb regeneration, Proc. Natl. Acad. Sci. U.S.A. 110 (2013) 9415–
9420.
[72] T.A. Petrie, N.S. Strand, C. Tsung-Yang, J.S. Rabinowitz, R.T. Moon,
Macrophages modulate adult zebrafish tail fin regeneration, Development
141 (2014) 2581–2591.
[73] A.B. Aurora, E.R. Porrello, W. Tan, A.I. Mahmoud, J.A. Hill, R. Bassel-Duby, H.A.
Sadek, E.N. Olson, Macrophages are required for neonatal heart regeneration,
J. Clin. Invest. 124 (2014) 1382–1392.
[74] V.E. Miron, A. Boyd, J.W. Zhao, T.J. Yuen, J.M. Ruckh, J.L. Shadrach, P. Van
Wijngaarden, A.J. Wagers, A. Williams, R.J. Franklin, C. Ffrench-Constant, M2
microglia and macrophages drive oligodendrocyte differentiation during CNS
remyelination, Nat. Neurosci. 16 (2013) 1211–1218.
[75] D. Dal-Secco, J. Wang, Z. Zeng, E. Kolaczkowska, C.H. Wong, B. Petri, R.M.
Ransohoff, I.F. Charo, C.N. Jenne, P. Kubes, A dynamic spectrum of monocytes
arising from the in situ reprogramming of CCR2+ monocytes at a site of sterile
injury, J. Exp. Med. 212 (2015) 447–456.
[76] P. Ramachandran, A. Pellicoro, M.A. Vernon, L. Boulter, R.L. Aucott, A. Ali, S.N.
Hartland, V.K. Snowdon, A. Cappon, T.T. Gordon-Walker, M.J. Williams, D.R.
Dunbar, J.R. Manning, N. van Rooijen, J.A. Fallowfield, S.J. Forbes, J.P. Iredale,
Differential Ly-6C expression identifies the recruited macrophage phenotype,
which orchestrates the regression of murine liver fibrosis, Proc. Natl. Acad.
Sci. U.S.A. 109 (2012) E3186–E3195.
[77] J.A. Stefater, S. Ren, R.A. Lang, J.S. Duffield, Metchnikoff’s policemen:
macrophages in development, homeostasis and regeneration, Trends Mol.
Med. 17 (2011) 743–752.
[78] K. Stark, A. Eckart, S. Haidari, A. Tirniceriu, M. Lorenz, M.L. von Bruhl, F.
Gartner, A.G. Khandoga, K.R. Legate, R. Pless, I. Hepper, K. Lauber, B. Walzog, S.
Massberg, Capillary and arteriolar pericytes attract innate leukocytes exiting
through venules and ’instruct’ them with pattern-recognition and motility
programs, Nat. Immunol. 14 (2013) 41–51.
Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28 25
[79] A. Armulik, G. Genove, C. Betsholtz, Pericytes: developmental, physiological,
and pathological perspectives, problems, and promises, Dev. Cell 21 (2011)
193–215.
[80] A. Dellavalle, G. Maroli, D. Covarello, E. Azzoni, A. Innocenzi, L. Perani, S.
Antonini, R. Sambasivan, S. Brunelli, S. Tajbakhsh, G. Cossu, Pericytes resident
in postnatal skeletal muscle differentiate into muscle fibres and generate
satellite cells, Nat. Commun. 2 (2011) 499.
[81] R. Katare, F. Riu, K. Mitchell, M. Gubernator, P. Campagnolo, Y. Cui, O.
Fortunato, E. Avolio, D. Cesselli, A.P. Beltrami, G. Angelini, C. Emanueli, P.
Madeddu, Transplantation of human pericyte progenitor cells improves the
repair of infarcted heart through activation of an angiogenic program
involving micro-RNA-132, Circ. Res. 109 (2011) 894–906.
[82] A. Pellicoro, P. Ramachandran, J.P. Iredale, J.A. Fallowfield, Liver fibrosis and
repair: immune regulation of wound healing in a solid organ, Nat. Rev.
Immunol. 14 (2014) 181–194.
[83] M. Vinish, W. Cui, E. Stafford, L. Bae, H. Hawkins, R. Cox, T. Toliver-Kinsky,
Dendritic cells modulate burn wound healing by enhancing early
proliferation, Wound Repair Regener. (2016).
[84] N. Gao, J. Yin, G.S. Yoon, Q.-S. Mi, X.Y. Fu-Shin, Dendritic cell-epithelium
interplay is a determinant factor for corneal epithelial wound repair, Am. J.
Pathol. 179 (2011) 2243–2253.
[85] A. Anzai, T. Anzai, S. Nagai, Y. Maekawa, K. Naito, H. Kaneko, Y. Sugano, T.
Takahashi, H. Abe, S. Mochizuki, Regulatory role of dendritic cells in
postinfarction healing and left ventricular remodeling, Circulation 125
(2012) 1234–1245.
[86] J. Jameson, K. Ugarte, N. Chen, P. Yachi, E. Fuchs, R. Boismenu, W.L. Havran, A
role for skin
c
dT cells in wound repair, Science 296 (2002) 747–749.
[87] T. Ono, K. Okamoto, T. Nakashima, T. Nitta, S. Hori, Y. Iwakura, H. Takayanagi,
IL-17-producing [gamma][delta] T cells enhance bone regeneration, Nat.
Commun. 7 (2016).
[88] K. Ramirez, D.A. Witherden, W.L. Havran, All hands on DE (T) C: Epithelial-
resident
c
dT cells respond to tissue injury, Cell. Immunol. 296 (2015) 57–61.
[89] P.L. Simonian, F. Wehrmann, C.L. Roark, W.K. Born, R.L. O’Brien, A.P. Fontenot,
C
dT cells protect against lung fibrosis via IL-22, J. Exp. Med. 207 (2010)
2239–2253.
[90] S. Reinke, S. Geissler, W. Taylor, K. Schmidt-Bleek, K. Juelke, V.
Schwachmeyer, Terminally differentiated CD8 (+) T cells negatively affect
bone regeneration in humans, Sci. Transl. Med. 5 (177) (2013) 177ra36,
http://dx.doi.org/10.1126/scitranslmed.3004754. Epub 2013/03/22, 5/177/
177ra36 [pii], PMID: 23515078.
[91] Y. Liu, L. Wang, T. Kikuiri, K. Akiyama, C. Chen, X. Xu, R. Yang, W. Chen, S.
Wang, S. Shi, Mesenchymal stem cell-based tissue regeneration is governed
by recipient T lymphocytes via IFN-[gamma] and TNF-[alpha], Nat. Med. 17
(2011) 1594–1601.
[92] D. Toben, I. Schroeder, T. El Khassawna, M. Mehta, J.-E. Hoffmann, J.-T. Frisch,
H. Schell, J. Lienau, A. Serra, A. Radbruch, G.N. Duda, Fracture healing is
accelerated in the absence of the adaptive immune system, J. Bone Miner. Res.
26 (2011) 113–124.
[93] A. Nosbaum, N. Prevel, H.-A. Truong, P. Mehta, M. Ettinger, T.C. Scharschmidt,
N.H. Ali, M.L. Pauli, A.K. Abbas, M.D. Rosenblum, Cutting edge: regulatory T
cells facilitate cutaneous wound healing, J. Immunol. 196 (2016) 2010–2014.
[94] H. Lei, K. Schmidt-Bleek, A. Dienelt, P. Reinke, H.-D. Volk, Regulatory T cell-
mediated anti-inflammatory effects promote successful tissue repair in both
indirect and direct manners, Front. Pharmacol. 6 (2015) 184.
[95] G. Trujillo, A.J. Hartigan, C.M. Hogaboam, T regulatory cells and attenuated
bleomycin-induced fibrosis in lungs of CCR7-/-mice, Fibrogenesis Tissue
Repair 3 (2010) 18.
[96] N.R. Aggarwal, F.R. D’Alessio, K. Tsushima, V.K. Sidhaye, C. Cheadle, D.N.
Grigoryev, K.C. Barnes, L.S. King, Regulatory T cell-mediated resolution of
lung injury: identification of potential target genes via expression profiling,
Physiol. Genomics 41 (2010) 109–119.
[97] B.T. Garibaldi, F.R. D’Alessio, J.R. Mock, D.C. Files, E. Chau, Y. Eto, M.B.
Drummond, N.R. Aggarwal, V. Sidhaye, L.S. King, Regulatory T cells reduce
acute lung injury fibroproliferation by decreasing fibrocyte recruitment, Am.
J. Respir. Cell Mol. Biol. 48 (2013) 35–43.
[98] L.-W. Lai, K.-C. Yong, Y.-H.H. Lien, Pharmacologic recruitment of regulatory T
cells as a novel therapy for ischemic acute kidney injury, Kidney Int. 81
(2012) 983–992.
[99] M.T. Gandolfo, H.R. Jang, S.M. Bagnasco, G.-J. Ko, P. Agreda, M.J. Soloski, M.T.
Crow, H. Rabb, Mycophenolate mofetil modifies kidney tubular injury and
Foxp3+ regulatory T cell trafficking during recovery from experimental
ischemia–reperfusion, Transpl. Immunol. 23 (2010) 45–52.
[100] D. Burzyn, W. Kuswanto, D. Kolodin, J.L. Shadrach, M. Cerletti, Y. Jang, E. Sefik,
T.G. Tan, A.J. Wagers, C. Benoist, A special population of regulatory T cells
potentiates muscle repair, Cell 155 (2013) 1282–1295.
[101] E. Rigamonti, P. Zordan, C. Sciorati, P. Rovere-Querini, S. Brunelli, Macrophage
plasticity in skeletal muscle repair, BioMed Res. Int. 2014 (2014).
[102] X. Meng, J. Yang, M. Dong, K. Zhang, E. Tu, Q. Gao, W. Chen, C. Zhang, Y. Zhang,
Regulatory T cells in cardiovascular diseases, Nat. Rev. Cardiol. 13 (2016)
167–179.
[103] G. Liu, H. Ma, L. Qiu, L. Li, Y. Cao, J. Ma, Y. Zhao, Phenotypic and functional
switch of macrophages induced by regulatory CD4+CD25+ T cells in mice,
Immunol. Cell Biol. 89 (2011) 130–142.
[104] M.W. Zaiss Dietmar, J. van Loosdregt, A. Gorlani, P.J. Bekker Cornelis, A. Gröne,
M.P. van Bergen en Henegouwen Paul, C. Roovers Rob, J. Coffer Paul, J.A.M.
Sijts Alice, Amphiregulin enhances regulatory T cell-suppressive function via
the epidermal growth factor receptor, Immunity 38 (2013) 275–284.
[105] N. Arpaia, J.A. Green, B. Moltedo, A. Arvey, S. Hemmers, S. Yuan, P.M. Treuting,
A.Y. Rudensky, A distinct function of regulatory T cells in tissue protection,
Cell 162 (2015) 1078–1089.
[106] W.L. Havran, J.M. Jameson, Epidermal T cells and wound healing, J. Immunol.
184 (2010) 5423–5428.
[107] L.L. Sharp, J.M. Jameson, G. Cauvi, W.L. Havran, Dendritic epidermal T cells
regulate skin homeostasis through local production of insulin-like growth
factor 1, Nat. Immunol. 6 (2005) 73–79.
[108] A. Toulon, L. Breton, K.R. Taylor, M. Tenenhaus, D. Bhavsar, C. Lanigan, R.
Rudolph, J. Jameson, W.L. Havran, A role for human skin–resident T cells in
wound healing, J. Exp. Med. 206 (2009) 743–750.
[109] P. Kumar, K. Rajasekaran, J.M. Palmer, M.S. Thakar, S. Malarkannan, IL-22: an
evolutionary missing-link authenticating the role of the immune system in
tissue regeneration, J. Cancer 4 (2013) 57–65.
[110] J. Kular, J. Tickner, S.M. Chim, J. Xu, An overview of the regulation of bone
remodelling at the cellular level, Clin. Biochem. 45 (2012) 863–873.
[111] I. Könnecke, A. Serra, T. El Khassawna, C. Schlundt, H. Schell, A. Hauser, A.
Ellinghaus, H.-D. Volk, A. Radbruch, G.N. Duda, T and B cells participate in
bone repair by infiltrating the fracture callus in a two-wave fashion, Bone 64
(2014) 155–165.
[112] R. Sridharan, A.R. Cameron, D.J. Kelly, C.J. Kearney, F.J. O’Brien, Biomaterial
based modulation of macrophage polarization: a review and suggested
design principles, Mater. Today 18 (2015) 313–325.
[113] S. Browne, A. Pandit, Biomaterial-mediated modification of the local
inflammatory environment, Front. Bioeng. Biotechnol. 3 (2015) 67.
[114] S.F. Badylak, J.E. Valentin, A.K. Ravindra, G.P. McCabe, A.M. Stewart-Akers,
Macrophage phenotype as a determinant of biologic scaffold remodeling,
Tissue Eng. Part A 14 (2008) 1835–1842.
[115] S.B. Orenstein, Y. Qiao, U. Klueh, D.L. Kreutzer, Y.W. Novitsky, Activation of
human mononuclear cells by porcine biologic meshes in vitro, Hernia: J.
Hernias Abdominal Wall Surgery 14 (2010) 401–407.
[116] J.A. Jones, D.T. Chang, H. Meyerson, E. Colton, I.K. Kwon, T. Matsuda, J.M.
Anderson, Proteomic analysis and quantification of cytokines and
chemokines from biomaterial surface-adherent macrophages and foreign
body giant cells, J. Biomed. Mater. Res., Part A 83 (2007) 585–596.
[117] F.Y. McWhorter, T. Wang, P. Nguyen, T. Chung, W.F. Liu, Modulation of
macrophage phenotype by cell shape, Proc. Natl. Acad. Sci. U.S.A. 110 (2013)
17253–17258.
[118] K. Nakamura, S. Yokohama, M. Yoneda, S. Okamoto, Y. Tamaki, T. Ito, M.
Okada, K. Aso, I. Makino, High, but not low, molecular weight hyaluronan
prevents T-cell-mediated liver injury by reducing proinflammatory cytokines
in mice, J. Gastroenterol. 39 (2004) 346–354.
[119] J.Y. Je, S.K. Kim, Reactive oxygen species scavenging activity of
aminoderivatized chitosan with different degree of deacetylation, Bioorg.
Med. Chem. 14 (2006) 5989–5994.
[120] N.J. Turner, S.F. Badylak, The use of biologic scaffolds in the treatment of
chronic nonhealing wounds, Adv. Wound Care 4 (2015) 490–500.
[121] S.F. Badylak, D.O. Freytes, T.W. Gilbert, Extracellular matrix as a
biological scaffold material: structure and function, Acta Biomater. 5
(2009) 1–13.
[122] B.N. Brown, B.D. Ratner, S.B. Goodman, S. Amar, S.F. Badylak, Macrophage
polarization: an opportunity for improved outcomes in biomaterials and
regenerative medicine, Biomaterials 33 (2012) 3792–3802.
[123] L. Huleihel, G.S. Hussey, J.D. Naranjo, L. Zhang, J.L. Dziki, N.J. Turner, D.B. Stolz,
S.F. Badylak, Matrix-bound nanovesicles within ECM bioscaffolds, Sci. Adv. 2
(2016) e1600502.
[124] A.J. Beattie, T.W. Gilbert, J.P. Guyot, A.J. Yates, S.F. Badylak, Chemoattraction
of progenitor cells by remodeling extracellular matrix scaffolds, Tissue Eng.
Part A 15 (2008) 1119–1125.
[125] K. Sadtler, K. Estrellas, B.W. Allen, M.T. Wolf, H. Fan, A.J. Tam, C.H.
Patel, B.S. Luber, H. Wang, K.R. Wagner, J.D. Powell, F. Housseau, D.M.
Pardoll, J.H. Elisseeff, Developing a pro-regenerative biomaterial scaffold
microenvironment requires T helper 2 cells, Science 352 (2016)
366–370.
[126] J.L. Dziki, D.S. Wang, C. Pineda, B.M. Sicari, T. Rausch, S.F. Badylak, Solubilized
extracellular matrix bioscaffolds derived from diverse source tissues
differentially influence macrophage phenotype, J. Biomed. Mater. Res., Part
A 105 (2017) 138–147.
[127] K.L. Spiller, S. Nassiri, C.E. Witherel, R.R. Anfang, J. Ng, K.R. Nakazawa, T. Yu, G.
Vunjak-Novakovic, Sequential delivery of immunomodulatory cytokines to
facilitate the M1-to-M2 transition of macrophages and enhance
vascularization of bone scaffolds, Biomaterials 37 (2015) 194–207.
[128] A. Asea, M. Rehli, E. Kabingu, J.A. Boch, O. Baré, P.E. Auron, M.A. Stevenson, S.
K. Calderwood, Novel signal transduction pathway utilized by extracellular
HSP70 role of Toll-like receptor (TLR) 2 and TLR4, J. Biol. Chem. 277 (2002)
15028–15034.
[129] J. Kovalchin, R. Wang, M. Wagh, J. Azoulay, M. Sanders, R. Chandawarkar, In
vivo delivery of heat shock protein 70 accelerates wound healing by up-
regulating macrophage-mediated phagocytosis, Wound Repair Regener. 14
(2006) 129–137.
[130] M. Yamamoto, T. Sato, J. Beren, D. Verthelyi, D.M. Klinman, The acceleration of
wound healing in primates by the local administration of immunostimulatory
CpG oligonucleotides, Biomaterials 32 (2011) 4238–4242.
26 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
[131] M. Kucia, K. Jankowski, R. Reca, M. Wysoczynski, L. Bandura, D.J. Allendorf, J.
Zhang, J. Ratajczak, M.Z. Ratajczak, CXCR4–SDF-1 signalling, locomotion,
chemotaxis and adhesion, J. Mol. Histol. 35 (2004) 233–245.
[132] T.T. Lau, D.-A. Wang, Stromal cell-derived factor-1 (SDF-1): homing factor for
engineered regenerative medicine, Exp. Opin. Biol. Ther. 11 (2011) 189–197.
[133] B.R. Son, L.A. Marquez-Curtis, M. Kucia, M. Wysoczynski, A.R. Turner, J.
Ratajczak, M.Z. Ratajczak, A. Janowska-Wieczorek, Migration of bone marrow
and cord blood mesenchymal stem cells in vitro is regulated by stromal-
derived factor-1-CXCR4 and hepatocyte growth factor-c-met axes and
involves matrix metalloproteinases, Stem Cells 24 (2006) 1254–1264.
[134] Y.H. Kim, Y. Tabata, Recruitment of mesenchymal stem cells and
macrophages by dual release of stromal cell-derived factor-1 and a
macrophage recruitment agent enhances wound closure, J. Biomed. Mater.
Res., Part A 104 (2016) 942–956.
[135] W. Shen, X. Chen, J. Chen, Z. Yin, B.C. Heng, W. Chen, H.-W. Ouyang, The effect
of incorporation of exogenous stromal cell-derived factor-1 alpha within a
knitted silk-collagen sponge scaffold on tendon regeneration, Biomaterials 31
(2010) 7239–7249.
[136] Y. Kimura, Y. Tabata, Controlled release of stromal-cell-derived factor-1 from
gelatin hydrogels enhances angiogenesis, J. Biomater. Sci. Polym. Ed. 21
(2010) 37–51.
[137] S.Y. Rabbany, J. Pastore, M. Yamamoto, T. Miller, S. Rafii, R. Aras, M. Penn,
Continuous delivery of stromal cell-derived factor-1 from alginate scaffolds
accelerates wound healing, Cell Transplant. 19 (2010) 399–408.
[138] G. Zhang, Y. Nakamura, X. Wang, Q. Hu, L.J. Suggs, J. Zhang, Controlled release
of stromal cell-derived factor-1alpha in situ increases c-kit+ cell homing to
the infarcted heart, Tissue Eng. 13 (2007) 2063–2071.
[139] P.T. Thevenot, A.M. Nair, J. Shen, P. Lotfi, C.-Y. Ko, L. Tang, The effect of
incorporation of SDF-1
a
into PLGA scaffolds on stem cell recruitment and the
inflammatory response, Biomaterials 31 (2010) 3997–4008.
[140] D. Projahn, S. Simsekyilmaz, S. Singh, I. Kanzler, B.K. Kramp, M. Langer, A.
Burlacu, J. Bernhagen, D. Klee, A. Zernecke, Controlled intramyocardial
release of engineered chemokines by biodegradable hydrogels as a
treatment approach of myocardial infarction, J. Cell Mol. Med. 18 (2014)
790–800.
[141] J. Ding, K. Hori, R. Zhang, Y. Marcoux, D. Honardoust, H.A. Shankowsky, E.E.
Tredget, Tredget E.E., Stromal cell-derived factor 1 (SDF-1) and its receptor
CXCR4 in the formation of postburn hypertrophic scar (HTS), Wound Repair
Regener. 19 (2011) 568–578.
[142] K. Noguchi, I. Ishikawa, The roles of cyclooxygenase-2 and prostaglandin E2
in periodontal disease, Periodontology 2000 (43) (2007) 85–101.
[143] L. Shapira, W.A. Soskolne, M. Sela, S. Offenbacher, V. Barak, The secretion of
PGE2, IL-1b, IL-6, and TNF
a
by adherent mononuclear cells from early onset
periodontitis patients, J. Periodontol. 65 (1994) 139–146.
[144] V. Paralkar, F. Borovecki, H. Ke, K. Cameron, B. Lefker, W. Grasser, T. Owen,
M. Li, P. DaSilva-Jardine, M. Zhou, An EP2 receptor-selective prostaglandin
E2 agonist induces bone healing, Proc. Natl. Acad. Sci. 100 (2003) 6736–
6740.
[145] S. Namkoong, S.-J. Lee, C.-K. Kim, Y.-M. Kim, H.-T. Chung, H. Lee, J.-A. Han, K.-
S. Ha, Y.-G. Kwon, Y.-M. Kim, Prostaglandin E2 stimulates angiogenesis by
activating the nitric oxide/cGMP pathway in human umbilical vein
endothelial cells, Exp. Mol. Med. 37 (2005) 588–600.
[146] P. Kalin
´ski, C. Hilkens, A. Snijders, F. Snijdewint, M.L. Kapsenberg, IL-12-
deficient dendritic cells, generated in the presence of prostaglandin E2,
promote type 2 cytokine production in maturing human naive T helper cells,
J. Immunol. 159 (1997) 28–35.
[147] L. Cui, S. Yin, W. Liu, N. Li, W. Zhang, Y. Cao, Expanded adipose-derived stem
cells suppress mixed lymphocyte reaction by secretion of prostaglandin E2,
Tissue Eng. 13 (2007) 1185–1195.
[148] K. Katamura, N. Shintaku, Y. Yamauchi, T. Fukui, Y. Ohshima, M. Mayumi, K.
Furusho, Prostaglandin E2 at priming of naive CD4+ T cells inhibits
acquisition of ability to produce IFN-gamma and IL-2, but not IL-4 and IL-5,
J. Immunol. 155 (1995) 4604–4612.
[149] P. Kamolratanakul, T. Hayata, Y. Ezura, A. Kawamata, C. Hayashi, Y.
Yamamoto, H. Hemmi, M. Nagao, R. Hanyu, T. Notomi, Nanogel-based
scaffold delivery of prostaglandin E2 receptor–specific agonist in
combination with a low dose of growth factor heals critical-size bone
defects in mice, Arthritis Rheum. 63 (2011) 1021–1033.
[150] N. Kato, U. Hasegawa, N. Morimoto, Y. Saita, K. Nakashima, Y. Ezura, H.
Kurosawa, K. Akiyoshi, M. Noda, Nanogel-based delivery system enhances
PGE2 effects on bone formation, J. Cell. Biochem. 101 (2007) 1063–1070.
[151] H. Toyoda, H. Terai, R. Sasaoka, K. Oda, K. Takaoka, Augmentation of bone
morphogenetic protein-induced bone mass by local delivery of a
prostaglandin E EP4 receptor agonist, Bone 37 (2005) 555–562.
[152] J.C. Sy, G. Seshadri, S.C. Yang, M. Brown, T. Oh, S. Dikalov, N. Murthy, M.E.
Davis, Sustained release of a p38 inhibitor from non-inflammatory
microspheres inhibits cardiac dysfunction, Nat. Mater. 7 (2008) 863–868.
[153] C.N. Serhan, Pro-resolving lipid mediators are leads for resolution physiology,
Nature 510 (2014) 92–101.
[154] C.N. Serhan, N. Chiang, J. Dalli, The resolution code of acute inflammation:
Novel pro-resolving lipid mediators in resolution, in: Semin. Immunol.,
Elsevier, 2015, pp. 200–215.
[155] A. Ariel, G. Fredman, Y.-P. Sun, A. Kantarci, T.E. Van Dyke, A.D. Luster, C.N.
Serhan, Apoptotic neutrophils and T cells sequester chemokines during
immune response resolution through modulation of CCR5 expression, Nat.
Immunol. 7 (2006) 1209–1216.
[156] Y. Tang, M.J. Zhang, J. Hellmann, M. Kosuri, A. Bhatnagar, M. Spite,
Proresolution therapy for the treatment of delayed healing of diabetic
wounds, Diabetes 62 (2013) 618–627.
[157] S. Hong, H. Tian, Y. Lu, J.M. Laborde, F.A. Muhale, Q. Wang, B.V. Alapure, C.N.
Serhan, N.G. Bazan, Neuroprotectin/protectin D1: endogenous biosynthesis
and actions on diabetic macrophages in promoting wound healing and
innervation impaired by diabetes, Am. J. Physiol. Cell Physiol. 307 (2014)
C1058–C1067.
[158] D.P. Vasconcelos, M. Costa, I.F. Amaral, M.A. Barbosa, A.P. Águas, J.N. Barbosa,
Modulation of the inflammatory response to chitosan through M2
macrophage polarization using pro-resolution mediators, Biomaterials 37
(2015) 116–123.
[159] D.P. Vasconcelos, A.C. Fonseca, M. Costa, I.F. Amaral, M.A. Barbosa, A.P. Águas,
J.N. Barbosa, Macrophage polarization following chitosan implantation,
Biomaterials 34 (2013) 9952–9959.
[160] C.P. De Oliveira, N. Magolbo, R. De Aquino, C. Weller, Oral aspirin for treating
venous leg ulcers, Cochrane Database Syst. Rev. 2 (2016). CD009432-CD.
[161] I. Cantón, R. Mckean, M. Charnley, K.A. Blackwood, C. Fiorica, A.J. Ryan, S.
MacNeil, Development of an Ibuprofen-releasing biodegradable PLA/PGA
electrospun scaffold for tissue regeneration, Biotechnol. Bioeng. 105 (2010)
396–408.
[162] L. Varatharajan, A. Thapar, T. Lane, A.B. Munster, A.H. Davies, Pharmacological
adjuncts for chronic venous ulcer healing: a systematic review, Phlebology
(2015). 0268355515587194.
[163] Q. Wang, H. Li, Y. Xiao, S. Li, B. Li, X. Zhao, L. Ye, B. Guo, X. Chen, Y. Ding,
Locally controlled delivery of TNF
a
antibody from a novel glucose-sensitive
scaffold enhances alveolar bone healing in diabetic conditions, J. Control.
Release 206 (2015) 232–242.
[164] E.E. Friedrich, L.T. Sun, S. Natesan, D.O. Zamora, R.J. Christy, N.R. Washburn,
Effects of hyaluronic acid conjugation on anti-TNF-
a
inhibition of
inflammation in burns, J. Biomed. Mater. Res., Part A 102 (2014) 1527–1536.
[165] Y.S. Kim, H.J. Park, M.H. Hong, P.M. Kang, J.P. Morgan, M.H. Jeong, J.G. Cho, J.C.
Park, Y. Ahn, TNF-alpha enhances engraftment of mesenchymal stem cells
into infarcted myocardium, Front. Biosci. 14 (2009) 2845–2856.
[166] W. Böcker, D. Docheva, W.C. Prall, V. Egea, E. Pappou, O. Roßmann, C. Popov,
W. Mutschler, C. Ries, M. Schieker, IKK-2 is required for TNF-
a
-induced
invasion and proliferation of human mesenchymal stem cells, J. Mol. Med. 86
(2008) 1183–1192.
[167] P.M. Mountziaris, A.G. Mikos, Modulation of the inflammatory response for
enhanced bone tissue regeneration, Tissue Eng. Part B: Rev. 14 (2008) 179–
186.
[168] C. Yan, W.A. Grimm, W.L. Garner, L. Qin, T. Travis, N. Tan, Y.-P. Han, Epithelial
to mesenchymal transition in human skin wound healing is induced by
tumor necrosis factor-
a
through bone morphogenic protein-2, Am. J. Pathol.
176 (2010) 2247–2258.
[169] Y. Ishida, T. Kondo, A. Kimura, K. Matsushima, N. Mukaida, Absence of IL-1
receptor antagonist impaired wound healing along with aberrant NF-
j
B
activation and a reciprocal suppression of TGF-bsignal pathway, J. Immunol.
176 (2006) 5598–5606.
[170] J. Chang, Z. Wang, E. Tang, Z. Fan, L. McCauley, R. Franceschi, K. Guan, P.H.
Krebsbach, C.-Y. Wang, Inhibition of osteoblastic bone formation by nuclear
factor-
j
B, Nat. Med. 15 (2009) 682–689.
[171] J. Chang, F. Liu, M. Lee, B. Wu, K. Ting, J.N. Zara, C. Soo, Hezaimi K. Al, W. Zou,
X. Chen, NF-
j
B inhibits osteogenic differentiation of mesenchymal stem cells
by promoting b-catenin degradation, Proc. Natl. Acad. Sci. 110 (2013) 9469–
9474.
[172] A. King, S. Balaji, L.D. Le, T.M. Crombleholme, S.G. Keswani, Regenerative
wound healing: the role of interleukin-10, Adv. Wound Care 3 (2014) 315–323.
[173] W.C. Chen, B.G. Lee, D.W. Park, K. Kim, H. Chu, K. Kim, J. Huard, Y. Wang,
Controlled dual delivery of fibroblast growth factor-2 and Interleukin-10 by
heparin-based coacervate synergistically enhances ischemic heart repair,
Biomaterials 72 (2015) 138–151.
[174] N. Mokarram, A. Merchant, V. Mukhatyar, G. Patel, R.V. Bellamkonda, Effect of
modulating macrophage phenotype on peripheral nerve repair, Biomaterials
33 (2012) 8793–8801.
[175] A.J. Rao, C. Nich, L.S. Dhulipala, E. Gibon, R. Valladares, S. Zwingenberger, R.L.
Smith, S.B. Goodman, Local effect of IL-4 delivery on polyethylene particle
induced osteolysis in the murine calvarium, J. Biomed. Mater. Res., Part A 101
(2013) 1926–1934.
[176] J.W. Penn, A.O. Grobbelaar, K.J. Rolfe, The role of the TGF-beta family in
wound healing, burns and scarring: a review, Int. J. Burns Trauma 2 (2012)
18–28.
[177] C.J. Johnston, D.J. Smyth, D.W. Dresser, R.M. Maizels, TGF-beta in tolerance,
development and regulation of immunity, Cell. Immunol. 299 (2016) 14–22.
[178] M.W. Ferguson, J. Duncan, J. Bond, J. Bush, P. Durani, K. So, L. Taylor, J.
Chantrey, T. Mason, G. James, Prophylactic administration of avotermin for
improvement of skin scarring: three double-blind, placebo-controlled, phase
I/II studies, Lancet 373 (2009) 1264–1274.
[179] P.M. Mountziaris, S.N. Tzouanas, D.C. Sing, P.R. Kramer, F.K. Kasper, A.G.
Mikos, Intra-articular controlled release of anti-inflammatory siRNA with
biodegradable polymer microparticles ameliorates temporomandibular joint
inflammation, Acta Biomater. 8 (2012) 3552–3560.
[180] F. Leuschner, P. Dutta, R. Gorbatov, T.I. Novobrantseva, J.S. Donahoe, G.
Courties, K.M. Lee, J.I. Kim, J.F. Markmann, B. Marinelli, P. Panizzi, W.W. Lee,
Y. Iwamoto, S. Milstein, H. Epstein-Barash, W. Cantley, J. Wong, V. Cortez-
Retamozo, A. Newton, K. Love, P. Libby, M.J. Pittet, F.K. Swirski, V. Koteliansky,
Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28 27
R. Langer, R. Weissleder, D.G. Anderson, M. Nahrendorf, Therapeutic siRNA
silencing in inflammatory monocytes in mice, Nat. Biotechnol. 29 (2011)
1005–1010.
[181] H.B. Sager, P. Dutta, J.E. Dahlman, M. Hulsmans, G. Courties, Y. Sun, T. Heidt, C.
Vinegoni, A. Borodovsky, K. Fitzgerald, G.R. Wojtkiewicz, Y. Iwamoto, B. Tricot,
O.F. Khan, K.J. Kauffman, Y. Xing, T.E. Shaw, P. Libby, R. Langer, R. Weissleder, F.
K. Swirski, D.G. Anderson, M. Nahrendorf, RNAi targeting multiple cell
adhesion molecules reduces immune cell recruitment and vascular
inflammation after myocardial infarction, Sci. Transl. Med. 8 (2016) 342ra80.
[182] H. Yin, R.L. Kanasty, A.A. Eltoukhy, A.J. Vegas, J.R. Dorkin, D.G. Anderson, Non-
viral vectors for gene-based therapy, Nat. Rev. Genet. 15 (2014) 541–555.
[183] A. Mehta, D. Baltimore, MicroRNAs as regulatory elements in immune system
logic, Nat. Rev. Immunol. 16 (2016) 279–294.
[184] A. Das, M. Sinha, S. Datta, M. Abas, S. Chaffee, C.K. Sen, S. Roy, Monocyte and
macrophage plasticity in tissue repair and regeneration, Am. J. Pathol. 185
(2015) 2596–2606.
[185] R.M. O’Connell, D.S. Rao, A.A. Chaudhuri, D. Baltimore, Physiological and
pathological roles for microRNAs in the immune system, Nat. Rev. Immunol.
10 (2010) 111–122.
[186] J. Banerjee, Y.C. Chan, C.K. Sen, MicroRNAs in skin and wound healing,
Physiol. Genomics 43 (2011) 543–556.
[187] J. Banerjee, C.K. Sen, MicroRNAs in skin and wound healing, MicroRNA
Protocols (2013) 343–356.
[188] D. Li, A. Wang, X. Liu, F. Meisgen, J. Grünler, I.R. Botusan, S. Narayanan, E.
Erikci, X. Li, L. Blomqvist, MicroRNA-132 enhances transition from
inflammation to proliferation during wound healing, J. Clin. Investig. 125
(2015) 3008–3026.
[189] Y. Singh, O.A. Garden, F. Lang, B.S. Cobb, MicroRNA-15b/16 enhances the
induction of regulatory T cells by regulating the expression of rictor and
mTOR, J. Immunol. 195 (2015) 5667–5677.
[190] D. Baumjohann, K.M. Ansel, MicroRNA-mediated regulation of T helper cell
differentiation and plasticity, Nat. Rev. Immunol. 13 (2013) 666–678.
[191] Q.-J. Li, J. Chau, P.J. Ebert, G. Sylvester, H. Min, G. Liu, R. Braich, M. Manoharan,
J. Soutschek, P. Skare, MiR-181a is an intrinsic modulator of T cell sensitivity
and selection, Cell 129 (2007) 147–161.
[192] A. Sethi, N. Kulkarni, S. Sonar, G. Lal, Role of miRNAs in CD4 T cell plasticity
during inflammation and tolerance, Front. Genet. 4 (2013) 8.
[193] J.E. Frith, E.R. Porrello, J.J. Cooper-White, Concise review: new frontiers in
microRNA-based tissue regeneration, Stem Cells Transl. Med. 3 (2014) 969–
976.
[194] T. Nakasa, M. Ishikawa, M. Shi, H. Shibuya, N. Adachi, M. Ochi, Acceleration of
muscle regeneration by local injection of muscle-specific microRNAs in rat
skeletal muscle injury model, J. Cell Mol. Med. 14 (2010) 2495–2505.
[195] Y. Zhang, Z. Wang, R.A. Gemeinhart, Progress in microRNA delivery, J. Control.
Release 172 (2013) 962–974.
[196] M. Gori, M. Trombetta, D. Santini, A. Rainer, Tissue engineering and
microRNAs: future perspectives in regenerative medicine, Exp. Opin. Biol.
Ther. 15 (2015) 1601–1622.
[197] J. Li, R. Kooger, M. He, X. Xiao, L. Zheng, Y. Zhang, A supramolecular hydrogel
as a carrier to deliver microRNA into the encapsulated cells, Chem. Commun.
50 (2014) 3722–3724.
[198] Y. Morishita, T. Imai, H. Yoshizawa, M. Watanabe, K. Ishibashi, S. Muto, D.
Nagata, Delivery of microRNA-146a with polyethylenimine nanoparticles
inhibits renal fibrosis in vivo, Int. J. Nanomed. 10 (2015) 3475–3488.
[199] E. van der Pol, A.N. Boing, P. Harrison, A. Sturk, R. Nieuwland, Classification,
functions, and clinical relevance of extracellular vesicles, Pharmacol. Rev. 64
(2012) 676–705.
[200] S.E.L. Andaloussi, I. Mager, X.O. Breakefield, M.J. Wood, Extracellular vesicles:
biology and emerging therapeutic opportunities, Nat. Rev. Drug Disc. 12
(2013) 347–357.
[201] A.M. Silva, J.H. Teixeira, M.I. Almeida, R.M. Gonçalves, M.A. Barbosa, S.G.
Santos, Extracellular vesicles: immunomodulatory messengers in the context
of tissue repair/regeneration, Eur. J. Pharm. Sci. (2016).
[202] J.H. Teixeira, A.M. Silva, M.I. Almeida, M.A. Barbosa, S.G. Santos, Circulating
extracellular vesicles: their role in tissue repair and regeneration, Transfusion
Apheresis Sci. 55 (2016) 53–61.
[203] S. Fais, L. O’Driscoll, F.E. Borras, E. Buzas, G. Camussi, F. Cappello, J. Carvalho,
A. Cordeiro da Silva, H. Del Portillo, S. El Andaloussi, Trcek T. Ficko, R. Furlan,
A. Hendrix, I. Gursel, V. Kralj-Iglic, B. Kaeffer, M. Kosanovic, M.E. Lekka, G.
Lipps, M. Logozzi, A. Marcilla, M. Sammar, A. Llorente, I. Nazarenko, C.
Oliveira, G. Pocsfalvi, L. Rajendran, G. Raposo, E. Rohde, P. Siljander, G. van
Niel, M.H. Vasconcelos, M. Yanez-Mo, M.L. Yliperttula, N. Zarovni, A.B. Zavec,
B. Giebel, Evidence-based clinical use of nanoscale extracellular vesicles in
nanomedicine, ACS Nano 10 (2016) 3886–3899.
[204] A. Marote, F.G. Teixeira, B. Mendes-Pinheiro, A.J. Salgado, MSCs-derived
exosomes: cell-secreted nanovesicles with regenerative potential, Front.
Pharmacol. 7 (2016) 231.
[205] T. Li, Y. Yan, B. Wang, H. Qian, X. Zhang, L. Shen, M. Wang, Y. Zhou, W. Zhu, W.
Li, W. Xu, Exosomes derived from human umbilical cord mesenchymal stem
cells alleviate liver fibrosis, Stem Cells Dev. 22 (2013) 845–854.
[206] J. Hyun, S. Wang, J. Kim, G.J. Kim, Y. Jung, MicroRNA125b-mediated Hedgehog
signaling influences liver regeneration by chorionic plate-derived
mesenchymal stem cells, Sci. Rep. 5 (2015) 14135.
[207] D.G. Phinney, M. Di Giuseppe, J. Njah, E. Sala, S. Shiva, C.M. St Croix, D.B. Stolz,
S.C. Watkins, Y.P. Di, G.D. Leikauf, J. Kolls, D.W. Riches, G. Deiuliis, N.
Kaminski, S.V. Boregowda, D.H. McKenna, L.A. Ortiz, Mesenchymal stem cells
use extracellular vesicles to outsource mitophagy and shuttle microRNAs,
Nat. Commun. 6 (2015) 8472.
[208] C. Lee, S.A. Mitsialis, M. Aslam, S.H. Vitali, E. Vergadi, G. Konstantinou, K.
Sdrimas, A. Fernandez-Gonzalez, S. Kourembanas, Exosomes mediate the
cytoprotective action of mesenchymal stromal cells on hypoxia-induced
pulmonary hypertension, Circulation 126 (2012) 2601–2611.
[209] Y. Qin, L. Wang, Z. Gao, G. Chen, C. Zhang, Bone marrow stromal/stem cell-
derived extracellular vesicles regulate osteoblast activity and differentiation
in vitro and promote bone regeneration in vivo, Sci. Rep. 6 (2016) 21961.
[210] L. Alvarez-Erviti, Y. Seow, H. Yin, C. Betts, S. Lakhal, M.J. Wood, Delivery of
siRNA to the mouse brain by systemic injection of targeted exosomes, Nat.
Biotechnol. 29 (2011) 341–345.
[211] M.M. Alvarez, J.C. Liu, Santiago G. Trujillo-de, B.-H. Cha, A. Vishwakarma, A.
Ghaemmaghami, A. Khademhosseini, Delivery strategies to control
inflammatory response: Modulating M1–M2 polarization in tissue
engineering applications, J. Control. Release (2016).
[212] Y.-H. Kim, H. Furuya, Y. Tabata, Enhancement of bone regeneration by dual
release of a macrophage recruitment agent and platelet-rich plasma from
gelatin hydrogels, Biomaterials 35 (2014) 214–224.
[213] A.J. Glowacki, R. Gottardi, S. Yoshizawa, F. Cavalla, G.P. Garlet, C. Sfeir, S.R.
Little, Strategies to direct the enrichment, expansion, and recruitment of
regulatory cells for the treatment of disease, Ann. Biomed. Eng. 43 (2015)
593–602.
[214] A.J. Glowacki, S. Yoshizawa, S. Jhunjhunwala, A.E. Vieira, G.P. Garlet, C. Sfeir,
S.R. Little, Prevention of inflammation-mediated bone loss in murine and
canine periodontal disease via recruitment of regulatory lymphocytes, Proc.
Natl. Acad. Sci. 110 (2013) 18525–18530.
[215] J. Nikolich-Zugich, Aging of the T cell compartment in mice and humans:
from no naive expectations to foggy memories, J. Immunol. 193 (2014) 2622–
2629.
28 Z. Julier et al. / Acta Biomaterialia 53 (2017) 13–28
... This potential role can also be applied to the wound healing process (Tonnesen et al., 2000), vascular growth during tissue regeneration (Saberianpour et al., 2018), and the context of ischemia (Hayashi et al., 2006). Furthermore, the activation of innate immune responses by hypoxic hAMSCs may play a crucial role in orchestrating the resolution of such pathologies (Julier et al., 2017). In contrast, the use of hAMSCs primed with IFN-γ, which has major immunosuppressive effects, might be therapeutically useful for the treatment of diseases characterized by an exacerbation of immune system activity (Chen and Brosnan, 2006;Leite et al., 2021;Kadri et al., 2023) (Figure 8). ...
Article
Full-text available
Mesenchymal stromal/stem cells (MSCs) are a heterogeneous population of multipotent cells that can be obtained from various tissues, such as dental pulp, adipose tissue, bone marrow and placenta. MSCs have gained importance in the field of regenerative medicine because of their promising role in cell therapy and their regulatory abilities in tissue repair and regeneration. However, a better characterization of these cells and their products is necessary to further potentiate their clinical application. In this study, we used unbiased high-resolution mass spectrometry-based proteomic analysis to investigate the impact of distinct priming strategies, such as hypoxia and IFN-γ treatment, on the composition and therapeutic functionality of the secretome produced by MSCs derived from the amniotic membrane of the human placenta (hAMSCs). Our investigation revealed that both types of priming improved the therapeutic efficacy of hAMSCs, and these improvements were related to the secretion of functional factors present in the conditioned medium (CM) and exosomes (EXOs), which play crucial roles in mediating the paracrine effects of MSCs. In particular, hypoxia was able to induce a pro-angiogenic, innate immune response-activating, and tissue-regenerative hAMSC phenotype, as highlighted by the elevated production of regulatory factors such as VEGFA, PDGFRB, ANGPTL4, ENG, GRO-γ, IL8, and GRO-α. IFN-γ priming, instead, led to an immunosuppressive profile in hAMSCs, as indicated by increased levels of TGFB1, ANXA1, THBS1, HOMER2, GRN, TOLLIP and MCP-1. Functional assays validated the increased angiogenic properties of hypoxic hAMSCs and the enhanced immunosuppressive activity of IFN-γ-treated hAMSCs. This study extends beyond the direct priming effects on hAMSCs, demonstrating that hypoxia and IFN-γ can influence the functional characteristics of hAMSC-derived secretomes, which, in turn, orchestrate the production of functional factors by peripheral blood cells. This research provides valuable insights into the optimization of MSC-based therapies by systematically assessing and comparing the priming type-specific functional features of hAMSCs. These findings highlight new strategies for enhancing the therapeutic efficacy of MSCs, particularly in the context of multifactorial diseases, paving the way for the use of hAMSC-derived products in clinical practice.
... (1) Advanced immunity One evolutionary innovation that could hamper the regenerative abilities of amniotes is their advanced immune system compared to that of anamniotes (Julier et al., 2017). The initial step of the immune response, the inflammatory Biological Reviews (2024) 000-000 © 2024 Cambridge Philosophical Society. ...
Article
The ability to regenerate large body appendages is an ancestral trait of vertebrates, which varies across different animal groups. While anamniotes (fish and amphibians) commonly possess this ability, it is notably restricted in amniotes (reptiles, birds, and mammals). In this review, we explore the factors contributing to the loss of regenerative capabilities in amniotes. First, we analyse the potential negative impacts on appendage regeneration caused by four evolutionary innovations: advanced immunity, skin keratinization, whole-body endothermy, and increased body size. These innovations emerged as amniotes transitioned to terrestrial habitats and were correlated with a decline in regeneration capability. Second, we examine the role played by the loss of regeneration-related enhancers and genes initiated by these innovations in the fixation of an inability to regenerate body appendages at the genomic level. We propose that following the cessation of regenerative capacity, the loss of highly specific regeneration enhancers could represent an evolutionarily neutral event. Consequently, the loss of such enhancers might promptly follow the suppression of regeneration as a side effect of evolutionary innovations. By contrast, the loss of regeneration-related genes, due to their pleiotropic functions, would only take place if such loss was accompanied by additional evolutionary innovations that compensated for the loss of pleiotropic functions unrelated to regeneration, which would remain even after participation of these genes in regeneration was lost. Through a review of the literature, we provide evidence that, in many cases, the loss in amniotes of genes associated with body appendage regeneration in anamniotes was significantly delayed relative to the time when regenerative capability was lost. We hypothesise that this delay may be attributed to the necessity for evolutionary restructuring of developmental mechanisms to create conditions where the loss of these genes was a beneficial innovation for the organism. Experimental investigation of the downregulation of genes involved in the regeneration of body appendages in anamniotes but absent in amniotes offers a promising avenue to uncover evolutionary innovations that emerged from the loss of these genes. We propose that the vast majority of regeneration-related genes lost in amniotes (about 150 in humans) may be involved in regulating the early stages of limb and tail regeneration in anamniotes. Disruption of this stage, rather than the late stage, may not interfere with the mechanisms of limb and tail bud development during embryogenesis, as these mechanisms share similarities with those operating in the late stage of regeneration. Consequently, the most promising approach to restoring regeneration in humans may involve creating analogs of embryonic limb buds using stem cell-based tissue-engineering methods, followed by their transfer to the amputation stump. Due to the loss of many genes required specifically during the early stage of regeneration, this approach may be more effective than attempting to induce both early and late stages of regeneration directly in the stump itself.
... M2 macrophages display regenerative and immunoprotective properties and usually appear in the later stages of inflammation, following interleukin (IL-4, IL-13, and IL-10) and TGF-β signals. They are mainly involved in tissue healing, ECM deposition, and angiogenesis but they also suppress the immune response and further activation of T-cells through the secretion of anti-inflammatory cytokines [48]. ...
Article
Full-text available
Adipose-derived stem cells (ADSCs) are mesenchymal stem cells with a great potential for self-renewal and differentiation. Exosomes derived from ADSCs (ADSC-exos) can imitate their functions, carrying cargoes of bioactive molecules that may affect specific cellular targets and signaling processes. Recent evidence has shown that ADSC-exos can mediate tissue regeneration through the regulation of the inflammatory response, enhancement of cell proliferation, and induction of angiogenesis. At the same time, they may promote wound healing as well as the remodeling of the extracellular matrix. In combination with scaffolds, they present the future of cell-free therapies and promising adjuncts to reconstructive surgery with diverse tissue-specific functions and minimal adverse effects. In this review, we address the main characteristics and functional properties of ADSC-exos in tissue regeneration and explore their most recent clinical application in wound healing, musculoskeletal regeneration, dermatology, and plastic surgery as well as in tissue engineering.
... A popular hypothesis in regenerative biology suggests that adult mammals have developed a more robust adaptive immune response at the expense of regenerative capabilities [1]. However, NOD/SCID mice that exhibit T cell deficiency but retain macrophage numbers fail to regenerate their hearts at neonatal stages and demonstrate signs of severe fibrosis, suggesting that modulating adaptive immunity alone is not sufficient for successful regeneration [2]. ...
Article
Full-text available
Background Previous studies have suggested that macrophages are present during lens regeneration in newts, but their role in the process is yet to be elucidated. Methods Here we generated a transgenic reporter line using the newt, Pleurodeles waltl, that traces macrophages during lens regeneration. Furthermore, we assessed early changes in gene expression during lens regeneration using two newt species, Notophthalmus viridescens and Pleurodeles waltl. Finally, we used clodronate liposomes to deplete macrophages during lens regeneration in both species and tested the effect of a subsequent secondary injury after macrophage recovery. Results Macrophage depletion abrogated lens regeneration, induced the formation of scar-like tissue, led to inflammation, decreased iris pigment epithelial cell (iPEC) proliferation, and increased rates of apoptosis in the eye. Some of these phenotypes persisted throughout the last observation period of 100 days and could be attenuated by exogenous FGF2 administration. A distinct transcript profile encoding acute inflammatory effectors was established for the dorsal iris. Reinjury of the newt eye alleviated the effects of macrophage depletion, including the resolution of scar-like tissue, and re-initiated the regeneration process. Conclusions Together, our findings highlight the importance of macrophages for facilitating a pro-regenerative environment in the newt eye by regulating fibrotic responses, modulating the overall inflammatory landscape, and maintaining the proper balance of early proliferation and late apoptosis of the iPECs.
Article
Full-text available
Stem cell therapy holds promise for tissue regeneration, yet significant challenges persist. Emerging as a safer and potentially more effective alternative, extracellular vesicles (EVs) derived from stem cells exhibit remarkable abilities to activate critical signaling cascades, thereby facilitating tissue repair. EVs, nano-scale membrane vesicles, mediate intercellular communication by encapsulating a diverse cargo of proteins, lipids, and nucleic acids. Their therapeutic potential lies in delivering cargos, activating signaling pathways, and efficiently mitigating oxidative stress—an essential aspect of overcoming limitations in stem cell-based tissue repair. This review focuses on engineering and applying EVs in tissue regeneration, emphasizing their role in regulating reactive oxygen species (ROS) pathways. Additionally, we explore strategies to enhance EV therapeutic activity, including functionalization and incorporation of antioxidant defense proteins. Understanding these molecular mechanisms is crucial for optimizing EV-based regenerative therapies. Insights into EV and ROS signaling modulation pave the way for targeted and efficient regenerative therapies harnessing the potential of EVs.
Article
Full-text available
Tissue engineering has demonstrated its efficacy in promoting tissue regeneration, and extensive research has explored its application in rotator cuff (RC) tears. However, there remains a paucity of research translating from bench to clinic. A key challenge in RC repair is the healing of tendon–bone interface (TBI), for which bioactive materials suitable for interface repair are still lacking. The umbilical cord (UC), which serves as a vital repository of bioactive components in nature, is emerging as an important source of tissue engineering materials. A minimally manipulated approach is used to fabricate UC scaffolds that retain a wealth of bioactive components and cytokines. The scaffold demonstrates the ability to modulate the TBI healing microenvironment by facilitating cell proliferation, migration, suppressing inflammation, and inducing chondrogenic differentiation. This foundation sets the stage for in vivo validation and clinical translation. Following implantation of UC scaffolds in the canine model, comprehensive assessments, including MRI and histological analysis confirm their efficacy in inducing TBI reconstruction. Encouraging short‐term clinical results further suggest the ability of UC scaffolds to effectively enhance RC repair. This investigation explores the mechanisms underlying the promotion of TBI repair by UC scaffolds, providing key insights for clinical application and translational research.
Chapter
Hard-to-heal wounds are an important public health issue worldwide, with a significant impact on the quality of life of patients. It is estimated that approximately 1–2% of the global population suffers from difficult wounds, which can be caused by a variety of factors such as trauma, infections, chronic diseases like diabetes or obesity, or poor health conditions. Hard-to-heal wounds are often characterized by a slow and complicated healing process, which can lead to serious complications such as infections, pressure ulcers, scar tissue formation, and even amputations. These complications can have a significant impact on the mobility, autonomy, and quality of life of patients, leading to an increase in healthcare and social costs associated with wound care. The preparation of the wound bed is a key concept in the management of hard-to-heal wounds, with the aim of promoting an optimal environment for healing. The TIME (Tissue, Infection/Inflammation, Moisture, Edge) model is a systematic approach used to assess and manage wounds in a targeted and personalized way. The concept of TIMER, expanding the TIME model, further focuses on regenerative processes, paying particular attention to promoting tissue regeneration and wound healing in a more effective and comprehensive way. The new element introduced in the TIMER model is “Regeneration”, which highlights the importance of activating and supporting tissue regeneration processes to promote complete and lasting wound healing. Regenerative therapies can include a wide range of approaches, including cellular therapies, growth factors, bioactive biomaterials, stem cell therapies, and growth factor therapies. These therapies aim to promote the formation of new healthy tissues, reduce inflammation, improve vascularization, and stimulate cellular proliferation to accelerate wound closure and prevent complications. Thanks to continuous progress in research and development of regenerative therapies, more and more patients suffering from difficult wounds can benefit from innovative and promising solutions to promote faster and more effective healing, improve quality of life, and reduce the risk of long-term complications.
Article
Full-text available
The management and reconstruction of critical-sized segmental bone defects remain a major clinical challenge for orthopaedic clinicians and surgeons. In particular, regenerative medicine approaches that involve incorporating stem cells within tissue engineering scaffolds have great promise for fracture management. This narrative review focuses on the primary components of bone tissue engineering—stem cells, scaffolds, the microenvironment, and vascularisation—addressing current advances and translational and regulatory challenges in the current landscape of stem cell therapy for critical-sized bone defects. To comprehensively explore this research area and offer insights for future treatment options in orthopaedic surgery, we have examined the latest developments and advancements in bone tissue engineering, focusing on those of clinical relevance in recent years. Finally, we present a forward-looking perspective on using stem cells in bone tissue engineering for critical-sized segmental bone defects.
Article
Full-text available
In the dynamic landscape of tissue engineering, the integration of tissue-engineered constructs (TECs) faces a dual challenge—initiating beneficial inflammation for regeneration while avoiding the perils of prolonged immune activation. As TECs encounter the immediate reaction of the immune system upon implantation, the unique immunomodulatory properties of mesenchymal stem/stromal cells (MSCs) emerge as key navigators. Harnessing the paracrine effects of MSCs, researchers aim to craft a localized microenvironment that not only enhances TEC integration but also holds therapeutic promise for inflammatory-driven pathologies. This review unravels the latest advancements, applications, obstacles, and future prospects surrounding the strategic alliance between MSCs and TECs, shedding light on the immunological symphony that guides the course of regenerative medicine.
Article
Full-text available
Exosomes are membrane-enclosed nanovesicles (30–150 nm) that shuttle active cargoes between different cells. These tiny extracellular vesicles have been recently isolated from mesenchymal stem cells (MSCs) conditioned medium, a population of multipotent cells identified in several adult tissues. MSCs paracrine activity has been already shown to be the key mediator of their elicited regenerative effects. On the other hand, the individual contribution of MSCs-derived exosomes for these effects is only now being unraveled. The administration of MSCs-derived exosomes has been demonstrated to restore tissue function in multiple diseases/injury models and to induce beneficial in vitro effects, mainly mediated by exosomal-enclosed miRNAs. Additionally, the source and the culture conditions of MSCs have been shown to influence the regenerative responses induced by exosomes. Therefore, these studies reveal that MSCs-derived exosomes hold a great potential for cell-free therapies that are safer and easier to manipulate than cell-based products. Nevertheless, this is an emerging research field and hence, further studies are required to understand the full dimension of this complex intercellular communication system and how it can be optimized to take full advantage of its therapeutic effects. In this mini-review, we summarize the most significant new advances in the regenerative properties of MSCs-derived exosomes and discuss the molecular mechanisms underlying these effects.
Article
Full-text available
Biologic scaffold materials composed of extracellular matrix (ECM) have been used in a variety of surgical and tissue engineering/regenerative medicine applications and are associated with favorable constructive remodeling properties including angiogenesis, stem cell recruitment, and modulation of macrophage phenotype toward an anti-inflammatory effector cell type. However, the mechanisms by which these events are mediated are largely unknown. Matrix-bound nanovesicles (MBVs) are identified as an integral and functional component of ECM bioscaffolds. Extracellular vesicles (EVs) are potent vehicles of intercellular communication due to their ability to transfer RNA, proteins, enzymes, and lipids, thereby affecting physiologic and pathologic processes. Formerly identified exclusively in biologic fluids, the presence of EVs within the ECM of connective tissue has not been reported. In both laboratory-produced and commercially available biologic scaffolds, MBVs can be separated from the matrix only after enzymatic digestion of the ECM scaffold material, a temporal sequence similar to the functional activity attributed to implanted bioscaffolds during and following their degradation when used in clinical applications. The present study shows that MBVs contain microRNA capable of exerting phenotypical and functional effects on macrophage activation and neuroblastoma cell differentiation. The identification of MBVs embedded within the ECM of biologic scaffolds provides mechanistic insights not only into the inductive properties of ECM bioscaffolds but also into the regulation of tissue homeostasis.
Article
Toll-like receptor 4 (TLR4) plays a key role in the initiation of innate immunity and in the regulation of adaptive immune responses. Using microarray analysis and PCR, TLR4 expression was observed to increase in experimental murine skin wounds at the early stages, suggesting a role in wound inflammation. Closure of excisional wounds was significantly delayed in TLR4 deficient (C3H/HeJ) as compared to wild type mice, and both IL-1β and IL-6 production were significantly lower in the wounds of TLR4 deficient mice. EGF also markedly decreased in the wound edge of epidermis in TLR4 deficient mice. In vitro studies confirmed that a wound stimulus induces TLR4 mRNA expression in primary normal human epidermal keratinocytes (NHEK). In vitro injury also induced the phosphorylation of p38 and JNK MAPK, as well as the expression of IL-1β and TNF-α by NHEK. Blockade of TLR4 abolished the phosphorylation of p38 and JNK MAPK, and blockade of TLR4 and/or p38/JNK abolished IL-1β production. The results strongly suggest that inflammatory cytokine production by injured NHEK is stimulated via the TLR4-p38 and JNK MAPK signaling pathway. Together, the current results provide evidence for a role of TLR4 at sites of injury, and suggest that TLR4 is an important regulator of wound inflammation.
Conference Paper
As one of the most important oil crops, the high-accuracy rapeseed digital model contributes to agronomic horticulture, crop-breeding and plant genetics research. How to build a three-dimensional (3D) plant model, especially a high-quality leaf model, has great significance to the research on the rapeseed plant growth status. In this paper, we present a three-dimensional oilseed rape plant reconstruction method based on point cloud data. Firstly, we provide a novel approach to reconstruct 3D rapeseed shape model by utilizing the acquired 3D point clouds, and the original point cloud acquired via 3D laser scanning should be preprocessed, the main idea of processing includes de-noising and simplifying the original point clouds. Then the meshing method and mesh optimization process have been carried out to attain a high-quality 3D rapeseed plant mesh model. Thus the rapeseed plants can be non-destructively monitored by measuring the reconstructed rapeseed plant models, and we can obtain the needed parameters such as the leaf area. Experiments are implemented on real datasets, and the results indicate that the accuracy of the leaf area values are highly closed to the real values, which are measured by conventional digital processing methodology.
Article
Inflammation is a complex and highly regulated biological process, crucial for a variety of functions in the human body, from host response against infectious agents to initiation of repair/regeneration of injured tissues. In the context of tissue repair, the action of different immune cell populations and their interplay with tissue specific cells, including stem cells, is still being uncovered. Extracellular Vesicles (EV) are small membrane vesicles secreted by cells in a controlled manner, which can act local and systemically. The ability of EV to influence tissue repair and regeneration has been proposed as a physiologically intelligent and targeted strategy of cell communication. Herein, the role of EV in tissue repair is reviewed, summarising first their contribution to the regulation of immune cell function, and discussing the implications for the resolution of inflammation during repair. Next, the impact of EV on cell proliferation and differentiation, and on extracellular matrix remodelling, key aspects of the subsequent phases of tissue repair, is addressed. Finally, EV-based therapies are discussed, focusing on the application of naturally produced EV, and the use of EV as delivery vehicles.
Article
The host response to biomaterials is a critical determinant of their success or failure in tissue-repair applications. Macrophages are among the first responders in the host response to biomaterials and have been shown to be predictors of downstream tissue remodeling events. Biomaterials composed of mammalian extracellular matrix (ECM) in particular have been shown to promote distinctive and constructive remodeling outcomes when compared to their synthetic counterparts, a property that has been largely attributed to their ability to modulate the host macrophage response. ECM bioscaffolds are prepared by decellularizing source tissues such as dermis and small intestinal submucosa. The differential ability of such scaffolds to influence macrophage behavior has not been determined. The present study determines the effects of ECM bioscaffolds derived from eight different source tissues upon macrophage surface marker expression, protein content, phagocytic capability, metabolism, and antimicrobial activity. The results show that macrophages exposed to small intestinal submucosa (SIS), urinary bladder matrix (UBM), brain ECM (bECM), esophageal ECM (eECM), and colonic ECM (coECM) express a predominant M2-like macrophage phenotype, which is pro-remodeling and anti-inflammatory (iNOS-/Fizz1+/CD206+). In contrast, macrophage exposure to dermal ECM resulted in a predominant M1-like, pro-inflammatory phenotype (iNOS+/Fizz1-/CD206-), whereas liver ECM (LECM) and skeletal muscle ECM (mECM) did not significantly change the expression of these markers. All solubilized ECM bioscaffold treatments resulted in an increased macrophage antimicrobial activity, but no differences were evident in macrophage phagocytic capabilities, and macrophage metabolism was decreased following exposure to UBM, bECM, mECM, coECM, and dECM. The present work could have important implications when considering the macrophage response following ECM implantation for site-appropriate tissue remodeling.
Article
Extracellular vesicles (EVs) have been a growing interest of the scientific community in recent years due to the wide possibilities of their evaluation as biomarkers of disease, and their potential to be used as therapeutic agents or vehicles. EVs that circulate in plasma carry proteins and nucleic acids, potentially to distant locations in the body where they can interfere with several cellular processes. To aid understanding of this rapidly evolving field, circulating EVs, including immune cell-derived ones, are reviewed here. Their cellular origins and described functions are discussed in a perspective of their contribution to regenerative processes. Different techniques for EV engineering and examples of their application are reviewed as a strong future direction of EV research. A summary of important aspects yet to be addressed ties up this review.
Article
Myocardial infarction (MI) leads to a systemic surge of vascular inflammation in mice and humans, resulting in secondary ischemic complications and high mortality. We show that, in ApoE(-/-) mice with coronary ligation, increased sympathetic tone up-regulates not only hematopoietic leukocyte production but also plaque endothelial expression of adhesion molecules. To counteract the resulting arterial leukocyte recruitment, we developed nanoparticle-based RNA interference (RNAi) that effectively silences five key adhesion molecules. Simultaneously encapsulating small interfering RNA (siRNA)-targeting intercellular cell adhesion molecules 1 and 2 (Icam1 and Icam2), vascular cell adhesion molecule 1 (Vcam1), and E- and P-selectins (Sele and Selp) into polymeric endothelial-avid nanoparticles reduced post-MI neutrophil and monocyte recruitment into atherosclerotic lesions and decreased matrix-degrading plaque protease activity. Five-gene combination RNAi also curtailed leukocyte recruitment to ischemic myocardium. Therefore, targeted multigene silencing may prevent complications after acute MI.
Article
MicroRNAs (miRNAs) are crucial post-transcriptional regulators of haematopoietic cell fate decisions. They act by negatively regulating the expression of key immune development genes, thus contributing important logic elements to the regulatory circuitry. Deletion studies have made it increasingly apparent that they confer robustness to immune cell development, especially under conditions of environmental stress such as infectious challenge and ageing. Aberrant expression of certain miRNAs can lead to pathological consequences, such as autoimmunity and haematological cancers. In this Review, we discuss the mechanisms by which several miRNAs influence immune development and buffer normal haematopoietic output, first at the level of haematopoietic stem cells, then in innate and adaptive immune cells. We then discuss the pathological consequences of dysregulation of these miRNAs.