ArticlePDF Available

Bone marrow stem cells have the ability to populate the entire central nervous system into fully differentiated parenchymal microglia

Wiley
The FASEB Journal
Authors:
  • Northern Ontario School of Medicine, Sudbury, Canada

Abstract and Figures

Pluripotent stem cells can differentiate into a variety of cell types during tissue development and regeneration. However, it is still unclear whether bone marrow-derived stem cells can migrate across the blood-brain barrier in many regions of the central nervous system (CNS) and if these cells can readily differentiate into functional parenchymal microglia. We thus studied the differentiation fate of bone marrow stem cells upon immigration into the CNS. To this end, we systemically transplanted stem cells that express green fluorescent protein (GFP) into lethally irradiated mice and found that these cells immigrated into the brain parenchyma of many regions of the CNS. Nearly all of the infiltrating cells had a highly ramified morphology and colocalized with the microglial marker iba1. Moreover, these cells expressed high levels of the protein CD11c, indicating that microglia of bone marrow origin may be potent antigen presenting cells. These data suggest that microglia of blood origin could activate cells of the adaptive immune system and cause harm to the CNS. Therefore, these results may have great clinical relevance for both immune-derived neuronal disorders and cancer patients undergoing allogeneic hematopoietic stem-cell transplantation.
Content may be subject to copyright.
The FASEB Journal express article 10.1096/fj.04-1517fje. Published online April 14, 2004.
Bone marrow stem cells have the ability to populate the
entire central nervous system into fully differentiated
parenchymal microglia
Alain R. Simard and Serge Rivest
Laboratory of Molecular Endocrinology, CHUL Research Center and Department of Anatomy
and Physiology, Laval University, Québec, Canada G1V 4G2
Corresponding author: S. Rivest, Laboratory of Molecular Endocrinology, CHUL Research
Center and Department of Anatomy and Physiology, Laval University, 2705 Laurier boul,
Québec, Canada G1V 4G2. E-mail: Serge.Rivest@crchul.ulaval.ca
ABSTRACT
Pluripotent stem cells can differentiate into a variety of cell types during tissue development and
regeneration. However, it is still unclear whether bone marrow-derived stem cells can migrate
across the blood-brain barrier in many regions of the central nervous system (CNS) and if these
cells can readily differentiate into functional parenchymal microglia. We thus studied the
differentiation fate of bone marrow stem cells upon immigration into the CNS. To this end, we
systemically transplanted stem cells that express green fluorescent protein (GFP) into lethally
irradiated mice and found that these cells immigrated into the brain parenchyma of many regions
of the CNS. Nearly all of the infiltrating cells had a highly ramified morphology and colocalized
with the microglial marker iba1. Moreover, these cells expressed high levels of the protein
CD11c, indicating that microglia of bone marrow origin may be potent antigen presenting cells.
These data suggest that microglia of blood origin could activate cells of the adaptive immune
system and cause harm to the CNS. Therefore, these results may have great clinical relevance for
both immune-derived neuronal disorders and cancer patients undergoing allogeneic
hematopoietic stem-cell transplantation.
Key words: Chimeric mice
GFP mice gliogenesis inflammation innate immunity antigen-
presenting cells
macrophages CD11c
t is a well-accepted fact that stem cells are capable of differentiating into multiple cell types
(1-3). In the central nervous system (CNS), however, neural stem cells have the ability to
give rise to astrocytes and oligodendrocytes but not microglia (4, 5). It has been suggested
that microglia are replenished partly by division of resident cells and partly by immigration of
circulating monocytes (6), though evidence for the latter was based on many assumptions. Two
previous studies have found many cells with microglial morphology associated with blood
vessels and were thus identified as perivascular microglial cells. However, both papers reported
that bone marrow-derived cells rarely crossed the blood-brain barrier (BBB) and
transdifferentiated into parenchymal microglia (7, 8). Moreover, this specific type of cell was
only consistently observed in the cerebellum (8). Therefore, the question still remains whether
I
Page 1 of 9
(page number not for citation purposes)
circulating stem cells can readily cross the BBB and differentiate into parenchymal microglia in
other regions of the brain.
To test this hypothesis, we transplanted GFP-expressing bone marrow cells into lethally
irradiated mice. We then evaluated the fate of the donor-derived cells in the CNS. Many of these
cells were found in the parenchyma throughout multiple regions of the brain, and we show that
the vast majority of these cells had differentiated into microglia since they were colabeled with
the microglial marker iba1. Most importantly, these donor-derived microglial cells are likely to
be effective antigen-presenting cells (APCs), because they contained high constitutive levels of
the surface molecule CD11c. These data may be of great value for understanding the etiology of
many neurodegenerative diseases and for cancer patients undergoing allogeneic hematopoietic
stem-cell transplantation.
EXPERIMENTAL PROCEDURES
Animals
Adult male C57BL/6J mice (~25 g; Jackson Laboratory, Bar harbor, ME) were acclimated to
standard laboratory conditions (14 h light, 10 h dark cycle; lights on at 0600 and off at 2000 h)
with free access to rodent chow and water. Hemizygous transgenic mice expressing GFP under
control of the chicken β-actin promoter and cytomegalovirus enhancer were initially obtained
from the same vendor (Jackson Laboratory). A colony was then established and maintained in a
C57BL/6J background. GFP mice were used as cell donors at 3-5 months of age. All protocols
were conducted according to the Canadian Council on Animal Care guidelines, as administered
by the Laval University Animal Welfare Committee.
Irradiation and bone marrow transplantation
Mice were exposed to 10 gray total-body irradiation using a cobalt-60 source (Theratron-780
model, MDS Nordion, Ottawa, ON, Canada). A few hours later, the animals were injected via a
tail vein with ~5 x 10
6
bone marrow cells freshly collected from GFP mice. The cells were
aseptically harvested by flushing femurs with Dulbecco’s PBS containing 2% fetal bovine serum
(DPBS-FBS). The samples were combined, centrifuged, and passed through a 25 gauge needle
and then filtered through a 40 µm nylon mesh. Recovered cells were resuspended in DPBS-FBS
at a concentration of 5 x 10
6
viable nucleated cells per 200 µl. Irradiated mice transplanted with
this suspension were housed in autoclaved cages and treated with antibiotics (0.2 mg
trimethorpine and 1 mg sulfamethoxazole/ml of drinking water given for 7 days before and 2 wk
after irradiation).
Seven weeks after transplantation, the mice were deeply anesthetized via an intraperitoneal
injection of a mixture of ketamine hydrochloride and xylazine and then were rapidly perfused
transcardially with 0.9% saline containing 10 U/ml of heparin, followed by 4%
paraformaldehyde in sodium phosphate buffer (pH 7.6 at 4°C). Brains were rapidly removed
from the skulls, postfixed for 2 h, and then placed in a solution containing 20% sucrose diluted in
4% paraformaldehyde buffer overnight at 4°C. The frozen brains were mounted on a microtome
(Reichert-Jung, Cambridge Instruments Company, Deerfield, IL), frozen with dry ice, and cut
into 25 µm coronal sections from the olfactory bulb to the end of the medulla. The slices were
Page 2 of 9
(page number not for citation purposes)
collected in a cold cryoprotectant solution (0.05 M sodium phosphate buffer, pH 7.3, 30%
ethylene glycol, 20% glycerol) and stored at -20°C.
Immunohistochemistry
Free-floating sections (25 µm thick) were incubated for 30 min in KPBS containing 4% goat
serum, 1% BSA, and 0.4% Triton X-100. With the use of the same buffer solution, the sections
were then incubated for 90 min in primary Ab (monoclonal rat anti-CD31, 1:1000, PharMingen,
San Diego, CA; polyclonal rabbit anti-ionized calcium binding adaptor molecule 1 (iba1),
1:1000, provided by Dr. Y. Imai, National Institute of Neuroscience, Japan; polyclonal Armenian
hamster anti-CD11c, 1:500, PharMingen) at room temperature. The sections were then rinsed 4 x
5 min in KPBS, followed by a 90 min incubation in fluorochrome-conjugated goat secondary Ab
(anti-rat Rhodamine Red-X, 1:625, Jackson ImmunoResearch Laboratories, Inc., West Grove,
PA; anti-rabbit Alexa 633, 1:625, Molecular Probes, Eugene, OR; anti-Armenian hamster
Rhodamine Red-X, 1:1000). Sections were then rinsed 4 x 5 min in KPBS, mounted onto
gelatin-coated slides and coverslipped with antifade medium composed of 96 mM Tris-Hcl, pH
8.0, 24% glycerol, 9.6% polyvinylalcohol, and 2.5% diazabicyclooctane (Sigma). Images were
obtained using a Fluoview confocal microscope (FV-500, Olympus Optical) with x20 and x60
oil objectives. Confocal images were acquired by sequential scanning using a two-frame Kalman
filter and a z-separation of 0.25 µm. The images were then processed to enhance contrast and
sharpness using Adobe Photoshop 7 (Adobe Systems) and were assembled using Adobe
Illustrator (Adobe Systems).
RESULTS
In this study, irradiated mice were transplanted with stem cells expressing GFP (9). We then
assessed the outcome of the donor-derived cells in the CNS of these lethally irradiated mice. We
found that 7 wk after the bone marrow transplant, many GFP-expressing cells were present in
various regions of the CNS, from the olfactory bulb to the end of the medulla (Fig. 1
). Moreover,
most of the donor-derived cells were not associated with blood vessels, indicating that they had
effectively crossed the BBB and infiltrated into the brain parenchyma. Regions that consistently
showed a great number of infiltrating GFP-positive cells included the anterior olfactory nucleus;
medial part (AOM; Fig. 1a
and b); piriform cortex (Fig. 1c and d); the lateral septal nucleus (Fig.
1e and f); the hypothalamus (Fig. 1g and h); the amygdaloid area (Fig. 1i and j); the S1 layer of
the cerebral cortex, trunk region (SiTr, Fig. 1k
and l); the hippocampus (Fig. 1m and n); the
substantia nigra (Fig. 1o
and p); the midbrain (Fig. 1q and r); the medulla (Fig. 1s and t); the
cerebellum (Fig. 1u
and v); and the area postrema (Fig. 1w and x). Though GFP-positive cells
were also present in other regions of the CNS, they were fewer in number compared with the
regions mentioned above (data not shown). Of interest is the fact that these cells were found
across the whole rostro-caudal brain in both neurogenic and non-neurogenic areas, which
indicates that bone marrow precursors have the ability to populate the cerebral tissue in a
nonspecific manner.
We observed that nearly all of the GFP-positive cells had a highly ramified phenotype,
resembling the morphology of parenchymal microglia. We thus assessed whether the ramified
cells expressed the microglial marker iba1 (10). As expected, all of the GFP-positive cells with a
ramified morphology were immunoreactive to iba1, providing direct evidence that donor stem
Page 3 of 9
(page number not for citation purposes)
cells had differentiated into parenchymal microglia (Fig. 2). In this regard, both GFP cells and
endogenous microglial cells exhibited similar immunoreactive signal levels for iba1. We also
tested the possibility that irradiation caused a massive neurodegeneration or death of other cell
types in the brain, which could account for the high cellular immigration and transdifferentiation
observed using our protocols. We thus used the fluorochrome Fluoro-Jade B (FJB), which is a
sensitive and reliable marker for the histochemical localization of neuronal degeneration (11,
12), and positive signals were not found in the brain of the animals included in this study (data
not shown). Therefore, neurodegeneration may not account for the phenomenon presented in this
study.
Macrophages are known to be very efficient APCs, and they express high levels of major
histocompatibility complex type II (MHC-II), whereas brain microglia are poor APCs and have
low surface levels of MHC-II. It has been shown that MHC-II expression is correlated with
another membrane-bound protein known as CD11c (13). Therefore, to determine if newly
differentiated microglia of donor stem cell origin are good APCs, we assessed the expression
levels of CD11c in tissues of the irradiated mice by immunohistochemistry (Fig. 3
). We found
that all GFP-positive cells colocalized with CD11c staining, whereas virtually none of the
residential microglia exhibited a positive signal for this protein. These data thus suggest that
donor-derived microglia may be proficient APCs.
DISCUSSION
Taken together, our data clearly show that stem cells originating from the bone marrow have the
capacity to migrate across the BBB and to transdifferentiate into microglia in vivo and that
generation of these cells during adulthood does not solely depend on the proliferation of resident
cells. To this date, evidence suggested that stem cells could not generate parenchymal microglia
(4, 5), except for two papers showing that only a few infiltrating cells displayed this phenotype
(7, 8). One of these papers shows that this phenomenon seems to consistently occur only in the
cerebellum (8). This report showed that only very small amounts of bone marrow-derived cells
(<0.3% of total donor-derived cells) had infiltrated past the BBB and differentiated into
parenchymal microglia (8). To the contrary, our data show that the large majority of GFP cells
were clearly not associated with blood vessels, although a small percentage of cells derived from
the donor had a perivascular nature. It is difficult to explain why bone marrow precursors were
able to fully differentiate into parenchymal microglia throughout the cerebral tissue here,
whereas previously analyzed mice exhibited only perivascular cells except for the cerebellum
(8). Mice of both studies were exposed to the same total radiation (10 gray), although it was
delivered via different protocols. Indeed, the animals in the previous study were irradiated with
two doses of 5 gray separated by a 3 h interval (8) for a total of 10 gray, whereas a single
radiation session was administered to the animals of the present study.
The fact that we now demonstrate that hematopoietic cells are capable of infiltrating and
generating microglia in various regions of the CNS has wide implications. Indeed, MHC-II is
expressed only on the surface of activated APCs, such as macrophages, dendritic cells, and
microglial cells in the brain. Expression of MHC-II is necessary for APCs to present a specific
foreign antigen to T cells. Parenchymal microglial cells display the same characteristics as
macrophages, but they are not recognized as being very potent APCs and their MHC-II levels are
low compared with other APCs. It is also known that effective APCs express high CD11c levels,
Page 4 of 9
(page number not for citation purposes)
whereas parenchymal microglia contain very small quantities of this protein (13). Our data show
that microglia of bone-marrow stem cell origin express high levels of CD11c, which is a strong
indication that these newly differentiated parenchymal microglia are likely to be effective APCs.
This could have a critical impact on brain disorders that have an immune etiology, such as
multiple sclerosis (MS). Indeed, macrophage-derived IL-12 stimulates the differentiation of a
subset of T lymphocytes (CD4+) into T helper 1 cells (Th-1) that produce interferon gamma
(IFNγ). These activated Th-1 cells are actually believed to play a critical role in MS, especially
during the demyelinating episodes. Immigrating microglia from blood circulation may therefore
contribute to the development of such immune processes in the cerebral environment.
Monocyte chemoattractant protein 1 (MCP-1) is likely to be essential in this recruitment of bone
marrow-derived cells, because the release of this chemokine by endothelial cells is the key
mechanism for the chemoattraction of cells of monocytic lineage. In this regard, MCP-1 and its
receptor CCR2 have been implicated in a number of inflammatory diseases and the report that
mice lacking CCR2 are resistant to experimental autoimmune encephalomyelitis (EAE), the
animal model of MS, provided decisive evidence that this chemokine has a leading role in the
cerebral infiltrating process (14). This concept is further supported by the total absence of
mononulear cell inflammatory infiltrates within the cerebral tissue of CCR2-deficient mice and
immunized with the myelin oligodendrocyte glycoprotein peptide 35-55 (14) and the fact that
monocyte recruitment to the CNS is a prerequisite step for the development of inflammatory
lesions in EAE (15). Differentiated macrophages into the brain parenchyma are believed to be
directly responsible for the demyelination and axonal dysfunction, as a consequence of myelin
sheath destruction by these activated phagocytes. The presence of these cells in the CNS of intact
animals may play a critical role in the fate that the innate immune system may have on
neurodegenerative disorders.
These data may also have a great clinical impact for cancer patients exposed to different levels of
chemotherapy and radiotherapy, particularly for those undergoing allogeneic hematopoietic
stem-cell transplantation. A similar immigration process of bone marrow precursors is expected
to take place in the brain of these patients, but the physiological relevance of such a phenomenon
has yet to be unraveled. Because these cells are more efficient APCs than endogenous microglia,
it is tempting to propose that recruitment of bone marrow precursors may be a critical step in
brain diseases with an immune etiology.
ACKNOWLEDGMENTS
The Canadian Institutes in Health Research (CIHR) supported this research. Alain Simard is
supported by a Ph.D. Studentship from the CIHR. Serge Rivest is a CIHR Scientist and holds a
Canadian Research Chair (Junior) in Neuroimmunology. The authors thank Dr. Y. Imai
(National Institute of Neuroscience, Kodaira, Japan) for the anti-ionized calcium binding adaptor
molecule 1 (iba1) antibody.
REFERENCES
1. Weissman, I. L. (2000) Translating stem and progenitor cell biology to the clinic: barriers
and opportunities. Science 287, 1442-1446
Page 5 of 9
(page number not for citation purposes)
2. Weissman, I. L. (2000) Stem cells: units of development, units of regeneration, and units in
evolution. Cell 100, 157-168
3. Fuchs, E., and Segre, J. A. (2000) Stem cells: a new lease on life. Cell 100, 143-155
4. Fricker, R. A., Carpenter, M. K., Winkler, C., Greco, C., Gates, M. A., and Bjorklund, A.
(1999) Site-specific migration and neuronal differentiation of human neural progenitor cells
after transplantation in the adult rat brain. J. Neurosci. 19, 5990-6005
5. Gage, F. H. (2000) Mammalian neural stem cells. Science 287, 1433-1438
6. Lawson, L. J., Perry, V. H., and Gordon, S. (1992) Turnover of resident microglia in the
normal adult mouse brain. Neuroscience 48, 405-415
7. Hickey, W. F., Vass, K., and Lassmann, H. (1992) Bone marrow-derived elements in the
central nervous system: an immunohistochemical and ultrastructural survey of rat chimeras.
J. Neuropathol. Exp. Neurol. 51, 246-256
8. Vallieres, L., and Sawchenko, P. E. (2003) Bone marrow-derived cells that populate the
adult mouse brain preserve their hematopoietic identity. J. Neurosci. 23, 5197-5207
9. Brazelton, T. R., Rossi, F. M., Keshet, G. I., and Blau, H. M. (2000) From marrow to brain:
expression of neuronal phenotypes in adult mice. Science 290, 1775-1779
10. Ito, D., Imai, Y., Ohsawa, K., Nakajima, K., Fukuuchi, Y., and Kohsaka, S. (1998)
Microglia-specific localisation of a novel calcium binding protein, Iba1. Brain Res Mol
Brain Res. 57, 1-9
11. Soulet, D., and Rivest, S. (2003) Polyamines play a critical role in the control of the innate
immune response in the mouse CNS. J. Cell Biol. 162, 257-268
12. Nadeau, S., and Rivest, S. (2003) Glucocorticoids play a fundamental role in protecting the
brain during innate immune response. J. Neurosci. 23, 5536-5544
13. Fischer, H. G., and Reichmann, G. (2001) Brain dendritic cells and macrophages/microglia
in central nervous system inflammation. J. Immunol. 166, 2717-2726
14. Izikson, L., Klein, R. S., Charo, I. F., Weiner, H. L., and Luster, A. D. (2000) Resistance to
experimental autoimmune encephalomyelitis in mice lacking the CC chemokine receptor
(CCR)2. J. Exp. Med.
192, 1075-1080
15. Tran, E. H., Hoekstra, K., van Rooijen, N., Dijkstra, C. D., and Owens, T. (1998) Immune
invasion of the central nervous system parenchyma and experimental allergic
encephalomyelitis, but not leukocyte extravasation from blood, are prevented in
macrophage-depleted mice. J. Immunol. 161, 3767-3775
Received January 21, 2004; accepted March 5, 2004.
Page 6 of 9
(page number not for citation purposes)
Fig. 1
Figure 1. Widespread distribution of bone marrow-derived cells throughout parenchymal elements of the central
nervous system. A primary antibody directed against CD31 labeled with rhodamine red-X-conjugated secondary antibody
was used to stain the endothelium of brain capillaries (red blood vessels). Several highly ramified GFP-positive cells were
identified in various areas of the brain, namely anterior olfactory nucleus, medial part (AOM; a and b), piriform cortex (c
and d), lateral septal nucleus (e and f), hypothalamus (g and h), amygdala (i and j), S1 cortex, trunk region (SiTr, k and l),
hippocampus (m and n), substantia nigra (o and p), midbrain (q and r), medulla (s and t), cerebellum (u and v), and area
postrema (w and x). Note that nearly all GFP-positive cells (green) in these regions had a ramified microglia-like
morphology and most of these cells were not associated with blood vessels (red). Scale = 50 µm.
Page 7 of 9
(page number not for citation purposes)
Fig. 2
Figure 2.
Bone marrow-derived cells transdifferentiate into microglial cells. CD31-positive cells were stained with
a rhodamine red-X-conjugated secondary antibody (red blood vessels), whereas a primary antibody directed against
iba1 was used to label microglia. They were then stained with an Alexa 633-conjugated secondary antibody (blue).
Donor-derived cells (green) with a ramified morphology were always colocalized with iba1 staining (blue) and rarely
colocalized with blood vessels (red). These results provide direct anatomical evidence that highly ramified cells are
parenchymal microglia, which originate from circulating donor stem cells. Scale bar = 25 µm.
Page 8 of 9
(page number not for citation purposes)
Fig. 3
Figure 3.
Microglial cells derived from bone marrow stem cells exhibit characteristics of antigen-presentating
cells (APCs). Effective APCs express high levels of the surface molecule CD11c, which was visualized by
immunohistochemistry using an anti-CD11c primary antibody and a rhodamine red-X-conjugated secondary antibody.
Virtually all GFP-positive cells were immunoreactive to CD11c (red). Arrows indicate highly localized cellular expression
of the surface antigen CD11c. These data suggest that infiltrating microglia derived from bone marrow precursors are
immunologically competent to present antigens to cells of the adaptive immunity. Scale bar = 10 µm.
Page 9 of 9
(page number not for citation purposes)
... The ability of bone marrow-derived cells to stably engraft into the central nervous system (CNS) was best demonstrated in animal models transplanted with green fluorescent protein (GFP)-labeled hematopoietic stem cells. [18][19][20] The engrafted GFP+ donor cells expressed the . CC-BY-ND 4.0 International license It is made available under a is the author/funder, who has granted medRxiv a license to display the preprint in perpetuity. ...
... doi: medRxiv preprint microglial marker Iba1 indicating they were BMDM. 18 CNS engraftment was shown to require cranial irradiation as a conditioning agent; subsequent studies, however, demonstrated that Busulfan-based conditioning as well as immune or genetic ablation of endogenous microglial cells can also promote CNS engraftment. [21][22][23][24] In fact, recent studies have reported near complete microglial replacement by using a combination of immune ablation and cranial irradiation in animal models. ...
... Most CNS engrafted donor-derived cells in animal models are thought to be BMDM based on the expression of Iba1. 18 To confirm that the donor cells we detected in our patient samples are in fact BMDM, we established a novel combined Iba1-IHC and XY-FISH assay. ...
Preprint
Full-text available
Hematopoietic stem cell transplantation is a well known treatment of hematologic malignancies wherein nascent stem cells provide a regenerating marrow and immunotherapy against the tumor. The progeny of hematopoietic stem cells also populate a wide spectrum of tissues, including the brain, as bone marrow derived macrophages similar to microglial cells. We developed a sensitive and novel combined IHC and XY FISH assay to detect, quantify and characterize donor cells in the cerebral cortex of 19 female allogeneic stem cell transplant patients. We show that the number of male donor cells ranged from 0.14-3.0% of total cells or 1.2-25% of microglial cells. Using tyramide based fluorescent IHC we found at least 80% of the donor cells express the microglial marker Iba1 consistent with being bone marrow derived macrophages. The percentage of donor cells was related to pretransplant conditioning; donor cells from radiation based myeloablative cases averaged 8.1% of microglial cells, while those from non-myeloablative cases averaged only 1.3%. The number of donor cells in patients conditioned with Busulfan or Treosulfan based myeloablation were similar to TBI based conditioning; donor cells averaged 6.8% of microglial cells. Notably, patients who received multiple transplants and those with the longest post-transplant survival had the highest level of donor engraftment, with donor cells averaging 16.3% of microglial cells. Our work represents the largest study characterizing bone marrow-derived macrophages in post-transplant patients. The efficiency of engraftment observed in our study warrants future research on microglial replacement as a therapeutic option for disorders of the central nervous system. Key Points We developed a sensitive Iba1 IHC and XY FISH co-stain assay to detect bone marrow derived donor cells in the brain of HSCT patients. Donor cells in the CNS express Iba1 and average 8% of microglia after myeloablative HSCT; double transplants increase engraftment up to 25%.
... Most studies utilized autologous bone marrow-derived stem cells [3,[12][13][14][19][20][21][22]26,30,32,46]. The use of these stem cells to mend neural tissue was described with encouraging results [52][53][54][55][56] and reportedly showed only motoric improvement [12]; in addition, Prasad observed a slight improvement in outcomes [26]. Lee observed considerable improvement in the early post-therapeutic period but no improvement at later stages [19]. ...
Article
Full-text available
Citation: Fauzi, A.A.; Thamrin, A.M.H.; Permana, A.T.; Ranuh, I.G.M.A.R.; Hidayati, H.B.; Hamdan, M.; Wahyuhadi, J.; Suroto, N.S.; Lestari, P.; Chandra, P.S. Comparison of the Administration Route of Stem Cell Therapy for Ischemic Stroke: A Systematic Review and Meta-Analysis of the Clinical Outcomes and Safety. J. Clin. Med. 2023, 12, 2735. https://doi. Abstract: Stem cell treatment is emerging as an appealing alternative for stroke patients, but there still needs to be an agreement on the protocols in place, including the route of administration. This systematic review aimed to assess the efficacy and safety of the administration routes of stem cell treatment for ischemic stroke. A systematic review was performed according to the Preferred Reporting Items for Systematic Reviews and Meta-Analysis (PRISMA) guidelines. A comprehensive literature search was undertaken using the PubMed, Scopus, and Cochrane databases. A total of 21 publications on stem cell therapy for ischemic stroke were included. Efficacy outcomes were measured using the National Institutes of Health Stroke Scale (NIHSS), the modified Rankin Scale (mRS), and the Barthel index (BI). Intracerebral administration showed a better outcome than other routes, but a greater number of adverse events followed due to its invasiveness. Adverse events were shown to be related to the natural history of stroke not to the treatment. However, further investigation is required, since studies have yet to compare the different administration methods directly.
... As the BBB and microglial cell maturation are established, the question arises as to how microglia are renewed in the adult brain. The participation of bone marrow-derived cells in the repopulation of microglial cell niches was proposed in various conditions, especially after bone marrow transplantation [123][124][125][126][127][128], and in diseases such as stroke [129], cerebral ischemia [130], bacterial meningitis [131], entorhinal cortex lesions [132], Parkinson's disease [133], Alzheimer's disease [134,135], multiple sclerosis [136], facial nerve axotomy and autoimmune encephalitis [137], scrapie [138], and brain and peripheral nerve injury [139][140][141]. During aging and the transition from plasticity to proinflammatory activation in primary neurodegeneration, the latest data also suggest that many metabolic byproducts and mitochondrial components can serve as damage-associated molecules, creating an extracellular gradient and accumulation of reactive oxygen species, which in turn propagate the inflammatory neurodegeneration [142,143]. ...
Article
Full-text available
Microglia belong to tissue-resident macrophages of the central nervous system (CNS), representing the primary innate immune cells. This cell type constitutes ~7% of non-neuronal cells in the mammalian brain and has a variety of biological roles integral to homeostasis and pathophysiology from the late embryonic to adult brain. Its unique identity that distinguishes its “glial” features from tissue-resident macrophages resides in the fact that once entering the CNS, it is perennially exposed to a unique environment following the formation of the blood–brain barrier. Additionally, tissue-resident macrophage progenies derive from various peripheral sites that exhibit hematopoietic potential, and this has resulted in interpretation issues surrounding their origin. Intensive research endeavors have intended to track microglial progenitors during development and disease. The current review provides a corpus of recent evidence in an attempt to disentangle the birthplace of microglia from the progenitor state and underlies the molecular elements that drive microgliogenesis. Furthermore, it caters towards tracking the lineage spatiotemporally during embryonic development and outlining microglial repopulation in the mature CNS. This collection of data can potentially shed light on the therapeutic potential of microglia for CNS perturbations across various levels of severity.
... Transplanted stem cells have the ability to enter the CNS and maintain their differentiation potential [42]. It is obvious that the BBB allows stem cells from the blood to enter the CNS region, where they may perform their functions. ...
Article
Full-text available
Autism spectrum disorder (ASD) is a consortium of developmental conditions. As scientists have not yet identified the exact underlying cause for these disorders, it is not easy to narrow down a singular therapy to propose a reliable cure. The preponderance of research suggests that stem-cell therapy improves aspects of outcome measure scales in patients with ASD; therefore, future studies should give us more confidence in the results. This overview considers the data that have emerged from the small set of published trials conducted using different approaches in stem-cell therapy for ASD, evaluates their results and proposes additional steps that could be taken if this field of endeavor is to be pursued further.
... Research conducted on a transgenic animal model with mutant human presenilin 1 and a chimeric mouse/human β-amyloid precursor protein (APP swe ) has demonstrated that blood-derived microglia can carry out more effective phagocytosis of amyloid deposits in comparison with resident microglia in vivo as well as in-vitro. Higher levels of antigen presentation in the formers' case could be the probable reason [140,141]. Recent studies have validated the therapeutic potential of injecting primary cultured rat microglia, bone-marrow-derived microglia (BMDM) [140], mouse bone-marrow-derived microglia-like cells (BMDML) [89,142], and peripheral blood-derived microglia-like (PBDML) cells [143] for the elimination of amyloid deposits by a directional migration and consequent, cell-specific phagocytosis [144]. We have ideated the potential promising role of TGF-β1 Table1 Cytoskeleton-associated genes regulated by TGFβ1 ...
Article
Full-text available
Microtubule-associated protein, Tau has been implicated in Alzheimer's disease for its detachment from microtubules and formation of insoluble intracellular aggregates within the neurons. Recent findings have suggested the expulsion of Tau seeds in the extracellular domain and their prion-like propagation between neurons. Transforming Growth Factor-β1 (TGF-β1) is a ubiquitously occurring cytokine reported to carry out immunomodulation and neuroprotection in the brain. TGF-β-mediated regulation occurs at the level of neuronal survival and differentiation, glial activation (astrocyte and microglia), amyloid production–distribution–clearance and neurofibrillary tangle formation, all of which contributes to Alzheimer's pathophysiology. Its role in the reorganization of cytoskeletal architecture and remodelling of extracellular matrix to facilitate cellular migration has been well-documented. Microglia are the resident immune sentinels of the brain responsible for surveying the local microenvironment, migrating towards the beacon of pertinent damage and phagocytosing the cellular debris or patho-protein deposits at the site of insult. Channelizing microglia to target extracellular Tau could be a good strategy to combat the prion-like transmission and seeding problem in Alzheimer's disease. The current review focuses on reaffirming the role of TGF-β1 signalling in Alzheimer’s pathology and cytoskeletal reorganization and considers utilizing the approach of TGF-β-triggered microglia-mediated targeting of extracellular patho-protein, Tau, as a possible potential strategy to combat Alzheimer's disease.
... Recently, it has been reported that blood-derived myeloid cells can cross the blood-brain barrier and differentiate into fully functional macrophages [22][23][24]. Blood-derived myeloid cells, recruited by central nervous system (CNS) damage, are considered microglial reinforcements of comparable functions and are accordingly termed "blood-derived macrophages" [25][26][27]. Activated microglia and blood-derived macrophages, often collectively referred to as CNS macrophages, adopt a variety of functional phenotypes that contribute to the progression of neurodegeneration as well as CNS repair and protection [28,29]. ...
Article
Full-text available
Alzheimer's disease (AD) is a heterogeneous disease with complex clinicopathological characteristics. To date, the role of m6A RNA methylation in monocyte-derived macrophages involved in the progression of AD is unknown. In our study, we found that methyltransferase-like 3 (METTL3) deficiency in monocyte-derived macrophages improved cognitive function in an amyloid beta (Aβ)-induced AD mouse model. The mechanistic study showed that that METTL3 ablation attenuated the m6A modification in DNA methyltransferase 3A (Dnmt3a) mRNAs and consequently impaired YTH N6-methyladenosine RNA binding protein 1 (YTHDF1)-mediated translation of DNMT3A. We identified that DNMT3A bound to the promoter region of alpha-tubulin acetyltransferase 1 (Atat1) and maintained its expression. METTL3 depletion resulted in the down-regulation of ATAT1, reduced acetylation of α-tubulin and subsequently enhanced migration of monocyte-derived macrophages and Aβ clearance, which led to the alleviated symptoms of AD. Collectively, our findings demonstrate that m6A methylation could be a promising target for the treatment of AD in the future.
Article
Glioblastoma is the most common primary malignant tumour of the central nervous system and remains uniformly and rapidly fatal. The tumour‐associated macrophage (TAM) compartment comprises brain‐resident microglia and bone marrow‐derived macrophages (BMDMs) recruited from the periphery. Immune‐suppressive and tumour‐supportive TAM cell states predominate in glioblastoma, and immunotherapies, which have achieved striking success in other solid tumours have consistently failed to improve survival in this ‘immune‐cold’ niche context. Hypoxic and necrotic regions in the tumour core are found to enrich, especially in anti‐inflammatory and immune‐suppressive TAM cell states. Microglia predominate at the invasive tumour margin and express pro‐inflammatory and interferon TAM cell signatures. Depletion of TAMs, or repolarisation towards a pro‐inflammatory state, are appealing therapeutic strategies and will depend on effective understanding and classification of TAM cell ontogeny and state based on new single‐cell and spatial multi‐omic in situ profiling. Here, we explore the application of these datasets to expand and refine TAM characterisation, to inform improved modelling approaches, and ultimately underpin the effective manipulation of function.
Article
Full-text available
This review examines the role of impaired amyloid-β clearance in the accumulation of amyloid-β in the brain and the periphery, which is closely associated with Alzheimer's disease (AD) and cerebral amyloid angiopathy (CAA). The molecular mechanism underlying amyloid-β accumulation is largely unknown, but recent evidence suggests that impaired amyloid-β clearance plays a critical role in its accumulation. The review provides an overview of recent research and proposes strategies for efficient amyloid-β clearance in both the brain and periphery. The clearance of amyloid-β can occur through enzymatic or non-enzymatic pathways in the brain, including neuronal and glial cells, blood-brain barrier, interstitial fluid bulk flow, perivascular drainage, and cerebrospinal fluid absorption-mediated pathways. In the periphery, various mechanisms, including peripheral organs, immunomodulation/immune cells, enzymes, amyloid-β-binding proteins, and amyloid-β-binding cells, are involved in amyloid-β clearance. Although recent findings have shed light on amyloid-β clearance in both regions, opportunities remain in areas where limited data is available. Therefore, future strategies that enhance amyloid-β clearance in the brain and/or periphery, either through central or peripheral clearance approaches or in combination, are highly encouraged. These strategies will provide new insight into the disease pathogenesis at the molecular level and explore new targets for inhibiting amyloid-β deposition, which is central to the pathogenesis of sporadic AD (amyloid-β in parenchyma) and CAA (amyloid-β in blood vessels).
Article
Hematopoietic stem cell transplantation is a well-known treatment of hematologic malignancies wherein nascent stem cells provide a regenerating marrow and immunotherapy against the tumor. The progeny of hematopoietic stem cells also populate a wide spectrum of tissues, including the brain, as bone marrow derived macrophages similar to microglial cells. We developed a sensitive and novel combined IHC and XY FISH assay to detect, quantify and characterize donor cells in the cerebral cortex of 19 female allogeneic stem cell transplant patients. We show that the number of male donor cells ranged from 0.14-3.0% of total cells or 1.2-25% of microglial cells. Using tyramide based fluorescent IHC we found at least 80% of the donor cells express the microglial marker IBA1 consistent with being bone marrow derived macrophages. The percentage of donor cells was related to pretransplant conditioning; donor cells from radiation based myeloablative cases averaged 8.1% of microglial cells, while those from non-myeloablative cases averaged only 1.3%. The number of donor cells in patients conditioned with Busulfan or Treosulfan based myeloablation were similar to TBI based conditioning; donor cells averaged 6.8% of microglial cells. Notably, patients who received multiple transplants and those with the longest post-transplant survival had the highest level of donor engraftment, with donor cells averaging 16.3% of microglial cells. Our work represents the largest study characterizing bone marrow-derived macrophages in post-transplant patients. The efficiency of engraftment observed in our study warrants future research on microglial replacement as a therapeutic option for disorders of the central nervous system.
Article
The C-X-C chemokine ligand (CXCL) 1 and its receptor C-X-C chemokine receptor (CXCR) 2 are widely expressed in the peripheral nervous systems (PNS) and central nervous systems (CNS) and are involved in the development of inflammation and pain after various nerve injuries. Once a nerve is damaged, it affects not only the neuron itself but also lesions elsewhere in its dominant site. After the CXCL1/CXCR2 axis is activated, multiple downstream pathways can be activated, such as c-Raf/MAPK/AP-1, p-PKC-μ/p-ILK/NLRP3, JAK2/STAT3, TAK1/NF-κB, etc. These pathways in turn mediate cellular motility state or cell migration. CXCR2 is expressed on the surface of neutrophils and monocytes/macrophages. These cells can be recruited to the lesion through the CXCL1/CXCR2 axis to participate in the inflammatory response. The expression of CXCR2 in neurons can activate some pathways in neurons through the CXCL1/CXCR2 axis, thereby causing damage to neurons. CXCR2 is also expressed in astrocytes, and when CXCR2 activated, it increases the number of astrocytes but impairs their function. Since inflammation can occur at almost any site of injury, elucidating the mechanism of CXCL1/CXCR2 axis' influence on inflammation may provide a favorable target for clinical treatment. Therefore, this article reviews the research progress of the CXCL1/CXCR2 axis in neurological diseases, aiming to provide a more meaningful theoretical basis for the treatment of neurological diseases.
Article
Full-text available
Recently it has been shown that mRNA of Iba1 (ionized calcium binding adaptor molecule 1), which was a novel calcium binding protein cDNA-cloned by our group, is specifically expressed in microglia in cultures of rat brain cells [Imai et al. Biophys. Biochem. Res. Commun., 224 (1996) 855–862]. In the present study, immunocytochemical and immunohistochemical examinations demonstrated that Iba1 protein is expressed in microglia alone both in cultured brain cells and in the brain, respectively. In a mixed cell culture of embryonic rat brain, immunocytochemically positive for Iba1 protein were the microglia but it was not detectable in neurons, astroglia, or oligodendroglia. Immunohistochemical staining of adult rat brain sections showed Iba1 protein to be specifically localised in ramified microglia. In addition, immunohistochemical staining and immunoblot analysis of activated microglia in the facial nucleus after facial nerve axotomy shows that expression of Iba1 protein was upregulated and peaked at 7 days. These results indicated that localisation of Iba1 protein is restricted to microglia both in vitro and in vivo, and that Iba1 protein plays a role in regulating the function of microglia, especially in the activated microglia.
Article
Full-text available
Organ-specific autoimmune diseases are characterized by infiltrates, including T lymphocytes and activated macrophages. Macrophages and secondarily activated tissue resident counterparts can both present Ag to and contribute to cytokine secretion by T lymphocytes. We have previously shown a crucial role of peripheral macrophages in experimental allergic encephalomyelitis (EAE), a Th1-mediated demyelinating disease that serves as a an animal model for multiple sclerosis (MS), by their depletion using mannosylated liposome-encapsulated dichloromethylene diphosphonate (Cl2MDP). Here we describe studies to investigate the mechanisms by which macrophages contribute to the lesion formation in EAE, by studying the effect of Cl2MDP-containing mannosylated liposomes (Cl2MDP-mnL) on adoptively transferred EAE in SJL/J mice. Adoptive transfer of EAE with myelin basic protein-reactive CD4+ T cells to SJL/J mice was abrogated by Cl2MDP-mnL treatment. CD4+ T cell and MHC II+ B220+ B cell extravasation from blood vessels and Th1 cytokine production were not inhibited. However, invasion of the central nervous system intraparenchymal tissues by lymphocytes, F4/80+, Mac-1+, and MOMA-1+ macrophages was almost completely blocked after treatment with Cl2MDP-mnL. Furthermore, in Cl2MDP-mnL-treated mice, the myelin sheaths appeared completely normal, whereas, in the control groups, marked demyelination occurred. Production of TNF-alpha and inducible nitric oxide synthase, both associated with macrophage/microglial activation, was inhibited. This intervention reveals a role for macrophages in regulating the invasion of autoreactive T cells and secondary glial recruitment that ordinarily lead to demyelinating pathology in EAE and multiple sclerosis.
Article
Full-text available
Neural progenitor cells obtained from the embryonic human forebrain were expanded up to 10(7)-fold in culture in the presence of epidermal growth factor, basic fibroblast growth factor, and leukemia inhibitory growth factor. When transplanted into neurogenic regions in the adult rat brain, the subventricular zone, and hippocampus, the in vitro propagated cells migrated specifically along the routes normally taken by the endogenous neuronal precursors: along the rostral migratory stream to the olfactory bulb and within the subgranular zone in the dentate gyrus, and exhibited site-specific neuronal differentiation in the granular and periglomerular layers of the bulb and in the dentate granular cell layer. The cells exhibited substantial migration also within the non-neurogenic region, the striatum, in a seemingly nondirected manner up to approximately 1-1.5 mm from the graft core, and showed differentiation into both neuronal and glial phenotypes. Only cells with glial-like features migrated over longer distances within the mature striatum, whereas the cells expressing neuronal phenotypes remained close to the implantation site. The ability of the human neural progenitors to respond in vivo to guidance cues and signals that can direct their differentiation along multiple phenotypic pathways suggests that they can provide a powerful and virtually unlimited source of cells for experimental and clinical transplantation.
Article
This report defines the bone marrow-derived elements found in the central nervous system of adult rat radiation chimeras. Four cell types were identified which bore the major histocompatibility (MHC) class I molecules of the donor rat strain thereby indicating a marrow origin. They were: meningeal macrophages, perivascular "microglial" cells, lymphocytes and rare cells with parenchymal microglial morphology. These cells were examined by immunohistochemical methods at the light microscopic and ultrastructural levels. Extended descriptions of the perivascular marrow-derived elements and the parenchymal microglial cells are presented. These latter two cell types, which exist in humans, have a significant role in neuroimmune processes and most probably function as the antigen-presenting cells in the central nervous system of mammals.
Article
We undertook this study to determine whether the microglia, the resident macrophages of the central nervous system, turn over in the steady-state. The turnover of brain macrophages would lend support to the "Trojan Horse" hypothesis of central nervous system infection, since one origin of replacement cells is the circulating monocyte pool. We combined the immunohistochemical detection of F4/80, a specific macrophage marker, with [3H]thymidine incorporation and autoradiography in normal adult mice. We could detect double-labelled cells in the brains of mice perfused 60 min after isotope administration. Such cells were few in number, randomly scattered throughout the brain and had the morphology of typical resident cells. The labelling index at this survival time was 0.052 +/- 0.003%. Thus resident microglia can synthesise DNA in situ. After longer survival times, we detected larger numbers of double-labelled cells. F4/80+ cells with resident morphology, mitotic figures, pairs of closely apposed (daughter) cells and cells with rounded macrophage-like morphology, all exhibited silver labelling. Twenty-four hours after isotope administration the labelling index was 0.192 +/- 0.052%. From morphologic evidence and comparison of labelling indices at different survival times, we concluded that: (i) resident microglia can synthesise DNA and go on to divide in situ; (ii) cells are recruited from the circulating monocyte pool through an intact blood-brain barrier and rapidly differentiate into resident microglia. We estimate that the two processes contribute almost equally to the steady-state turnover of resident microglia.
Article
Stem cells are the natural units of embryonic generation, and also adult regeneration, of a variety of tissues. Recently, the list of tissues that use the model of differentiation from stem to progenitor to mature cell has increased from blood to include a variety of tissues, including both central and peripheral nervous systems and skeletal muscle; it is also possible that all organs and tissues are derived from, and still contain, stem cells. Because the number and activities of stem cells and their progeny are homeostatically regulated, clinical stem cell transplantation could greatly add to the physician's armamentarium against degenerative diseases.