You are on page 1of 15

Bone 134 (2020) 115303

Contents lists available at ScienceDirect

Bone
journal homepage: www.elsevier.com/locate/bone

Full Length Article

miR-99a in bone homeostasis: Regulating osteogenic lineage commitment T


and osteoclast differentiation
Sara Reis Mouraa,b, Joao Paulo Brasa,b,c, Jaime Freitasa,b, Hugo Osórioa,d,e,
Mario Adolfo Barbosaa,b,c, Susana Gomes Santosa,b,c, Maria Ines Almeidaa,b,c,∗
a
i3S - Instituto de Investigação e Inovação em Saúde, Universidade do Porto, 4200-135 Porto, Portugal
b
INEB - Instituto de Engenharia Biomédica, Universidade do Porto, 4200-135 Porto, Portugal
c
ICBAS - Instituto de Ciências Biomédicas Abel Salazar, Universidade do Porto, 4050-313 Porto, Portugal
d
Ipatimup - Instituto de Patologia e Imunologia Molecular da Universidade do Porto, 4200-135 Porto, Portugal
e
FMUP – Faculdade de Medicina da Universidade do Porto, 4200-319 Porto, Portugal

ARTICLE INFO ABSTRACT

Keywords: Background: The tight coupling between osteoblasts and osteoclasts is essential to maintain bone homeostasis.
Cell differentiation Deregulation of this process leads to loss and deterioration of the bone tissue causing diseases, such as osteo-
Osteoporosis porosis. MicroRNAs are able to control bone-related mechanisms and have been explored as therapeutic tools. In
Bone remodeling this study, we investigated the potential of miR-99a-5p to modulate osteogenic differentiation, osteoclasto-
Osteoprotegerin
genesis, and the osteoblasts-osteoclasts crosstalk.
Gene therapy
Methods: To achieve this goal, human primary Mesenchymal Stem/Stromal Cells (MSC) were differentiated into
osteoblasts and adipocytes, and miR-99a-5p expression was evaluated by RT-qPCR. Knockdown and over-
expression experiments were conducted to modulate miR-99a-5p expression in MC3T3 cells. Cell proliferation
and cell death/apoptosis were evaluated by resazurin assay and flow cytometry, respectively. Proteomic analysis
was used to identify the miR-99a-5p regulatory network, and ELISA to evaluate OPG levels in the cell culture
supernatant. Conditioned media from MC3T3-transfected cells was used to culture RAW 264.7 cells and the
effect on osteoclast differentiation was assessed. Human primary monocytes were isolated to induce osteo-
clastogenesis and evaluate miR-99a-5p expression. Finally, levels of miR-99a-5p were modulated in RAW 264.7
cells to understand the impact on osteoclastogenesis.
Results: The results show that miR-99a-5p is significantly downregulated during the early stages of human
primary MSCs osteogenic differentiation and during MC3T3 osteogenic differentiation. On the other hand, miR-
99a-5p levels are increased during the initial stages of adipogenic differentiation. Inhibition of miR-99a-5p in
MC3T3 pre-osteoblastic cells promoted osteogenic differentiation, whereas its overexpression suppressed the
levels of osteogenic specific genes (Runx2 and Alpl), as well as mineralization, with no effect on proliferation or
apoptosis. Proteomic analysis of miR-99a-5p-transfected cells showed that numerous proteins known to be in-
volved in cell differentiation were altered, including osteogenic differentiation markers and extracellular matrix
proteins. While inhibition of miR-99a-5p increased the Tnfrsf11b (OPG encoding gene)/Tnfsf11 (RANKL en-
coding gene) mRNA expression ratio, in addition to increasing OPG secretion, miR-99a-5p overexpression re-
sulted in the opposite effect. The cell culture supernatant of miR-99a-5p-inhibited MC3T3 cells impaired the
osteoclastogenic potential of RAW 264.7 cells by decreasing the number of multinucleated cells and reducing the
expression of osteoclastogenic markers. Interestingly, miR-99a-5p expression is increased during osteoclasts
differentiation, both in human primary monocytes and RAW 264.7. These results show that miR-99a-5p per se is
a positive regulator of osteoclastogenic differentiation.
Conclusions: Globally, our findings show that miR-99a-5p inhibition promotes the commitment into osteogenic
differentiation, impairs osteoclastogenic differentiation, and control bone cells communication. Ultimately, it
supports miR-99a-5p as a target candidate for future miRNA-based therapies for bone diseases associated with
bone remodeling deregulation.


Corresponding author at: Rua Alfredo Allen, 208, 4200-135 Porto, Portugal.
E-mail address: ines.almeida@ineb.up.pt (M.I. Almeida).

https://doi.org/10.1016/j.bone.2020.115303
Received 1 October 2019; Received in revised form 4 February 2020; Accepted 25 February 2020
Available online 29 February 2020
8756-3282/ © 2020 Elsevier Inc. All rights reserved.
S.R. Moura, et al. Bone 134 (2020) 115303

1. Introduction throughput proteomics, evaluated modulation of OPG secretion, ex-


plored miR-99a-5p as a player in the crosstalk between osteoblasts and
Bone remodeling involves the coupling between new bone forma- osteoclasts, and lastly, investigated miR-99a-5p levels and its impact on
tion by osteoblasts and bone resorption by osteoclasts [1]. In healthy osteoclastogenesis.
conditions, this process is tightly regulated, particularly through the
RANKL/RANK/OPG signalling pathway [2,3]. However, bone re- 2. Materials and methods
modeling can be disrupted by intrinsic or extrinsic factors. An im-
balance in the bone remodeling process may lead to the development of 2.1. Human primary bone marrow-derived MSCs
bone disorders, such as osteoporosis, which is a systemic progressive
skeletal disease, with an increased vulnerability to bone fragility frac- Human primary MSCs (hMSCs) were isolated from the discarded
tures [4,5]. The formation and growth of bone tissue after a fracture, bone marrow of patients that underwent total hip arthroplasty or
whether as a consequence of a pathological disease or an injury, is a anterior cruciate ligament injury at Centro Hospitalar de São João
complex process that involves distinct cell types, including the re- (CHSJ, Porto, Portugal), after signing informed consent. These patients
cruitment of Mesenchymal Stem/stromal Cells (MSCs) to the injury site, did not suffer from known inflammatory diseases. Information about
which then differentiate into osteoblasts [6]. During bone regenera- the donors' sex, age, site of bone marrow collection and type of surgery
tion/repair, it is crucial for MSCs to commit into the osteogenic lineage is provided in Supplemental Table 1. The protocol was approved by
in detriment of adipogenic differentiation [7]. In this context, the CHSJ Ethics Committee for Health, and conforms to the declaration of
identification and dissection of novel regulatory players in MSCs dif- Helsinki. hMSCs from 6 donors were cultured in low-glucose Dulbecco's
ferentiation, and in the crosstalk with osteoclasts, is essential to better Modified Eagle's Medium (DMEM, Corning) supplemented with 10%
understand the molecular processes that contribute to bone homeostasis (v/v) fetal bovine serum (FBS, mesenchymal stem cell-qualified, Gibco)
and to bone regeneration/repair. and 1% (v/v) penicillin/streptomycin (P/S, Invitrogen) in a humidified
MicroRNAs (miRNAs) are small endogenous non-coding RNAs that atmosphere at 37 °C and 5% (v/v) CO2. hMSCs purity was confirmed by
regulate gene expression post-transcriptionally, through com- flow cytometry through the expression of specific surface antigens
plementary binding to RNA transcripts, leading to RNA degradation or CD73, CD90 and CD105 (positive markers) and lack of expression of
inhibition of translation [8,9]. Recently, several studies described HLA-DR, CD19, CD14, CD34 and CD45 (negative markers), according
miRNAs as key regulators of bone related processes [10–13]. For in- to the criteria described by the International Society for Cellular
stance, the local miRNA expression pattern is changed upon bone Therapy [25].
fractures in healthy animal models [14,15] and in human osteoporotic
fractures [16–18]. Also, as previously demonstrated by us, miRNA 2.2. Human primary monocyte isolation
transcriptome is systemically changed in a timely manner upon a cri-
tical fracture in the rat femur [19]. We and others have identified Human primary monocytes were isolated from buffy coats of 6
miRNAs as regulators of MSCs proliferation, differentiation and com- healthy blood donors kindly provided by CHSJ. The protocol was ap-
munication with other cell types involved in bone repair, such as en- proved by CHSJ Ethics Committee for Health, and conforms to the
dothelial and immune cells [11,20,21]. Recently, miRNA microarray declaration of Helsinki. RosetteSep human monocyte enrichment iso-
data analysis performed by us in the pre-osteoblast mouse cell line lation kit was used (StemCell Technologies), as previously described by
MC3T3, and further confirmed in human primary MSCs, showed that us [20]. Briefly, buffy coats were centrifuged at 1200g, for 30 min, at
miR-29b-3p, miR-29c-3p and miR-20a-5p were overexpressed, whereas room temperature (RT), without active acceleration or brake. The
miR-143-3p, miR-195-5p and miR-497-5p were downregulated during Peripheral Blood Mononuclear Cells (PBMCs) layer was collected, to-
osteogenic differentiation [11]. The microarray data also identified gether with some red blood cells necessary for the formation of im-
miR-99a-5p to be downregulated in MC3T3 cells after 7 days of culture munorosettes. Following the manufacturer's instructions, the Ro-
in osteogenic-inducing conditions. Recently, Tang et al. has validated setteSep human monocyte enrichment isolation kit was added to the
miR-99a-5p role in the osteogenic differentiation of mouse-derived cells and incubated during 20 min at RT in a horizontal shaker. Samples
MSCs and bone healing in a mice model [22]. Interestingly, a miR-99a were then 1:1 diluted in phosphate buffer saline 1× (PBS) supple-
family member, miR-100, was shown to be upregulated in bone tissue mented with 2% (v/v) FBS (Biowest), before being carefully laid over
and serum samples from osteoporotic patients compared with non-os- Histopaque®-1077 (Sigma-Aldrich). After centrifugation at 1200g, for
teoporotic patients [17]. Furthermore, the expression of this miR-99a 20 min, at RT, without acceleration and brake, a second gradient was
family member negatively correlate with bone mineral density, both in formed and the enriched monocyte layer (intermediate layer) was
plasma [23], and in bone tissue samples [24]. Therefore, dissecting the collected and washed three times with PBS 1× by centrifugation at
mechanisms underlying 300g during 20 min, before plating. Lastly, cells were resuspended in
miR-99a-5p in bone cells is essential to support its usage as a Roswell Park Memorial Institute (RPMI, Gibco) with 10% (v/v) FBS
therapeutic target for future clinical approaches. (Biowest) and 1% P/S (v/v) and counted using trypan blue dye (Sigma-
Currently, the miRbase database (v22) includes > 30,000 entries, Aldrich) exclusion assay. Monocyte purity was routinely assessed by
out of which, > 2500 correspond to human mature miRNA sequences flow cytometry analysis for positive CD14 marker (Immunotools), as
[23]. Hence, selecting which miRNA are candidates for future bone- previously described [20,26]. Cells labelled with matching isotype were
related clinical applications remains challenging. Envisioning future used as negative control. Monocyte population contained > 80% CD14
miRNA-based therapies for bone repair/regeneration in the context of positive cells (Supplemental Fig. 1).
fragility fractures, as a consequence of osteoporosis or other patholo-
gies, the ideal miRNA candidate should be able to promotes MSCs os- 2.3. Cell lines
teogenic differentiation in detriment of other lineages, while con-
comitantly decrease osteoclasts formation or activity. In this context, MC3T3-E1 pre-osteoblastic cells, widely used in osteoblast biology
we explored the role of miR-99a-5p as a potential therapeutic target studies [27], and RAW 264.7 monocytic cells, commonly used in os-
using in vitro tools. Specifically, we determined the involvement of teoclasts studies due to their osteoclastogenic potential [28], were
miR-99a-5p in MSC lineage commitment by analysing it expression maintained in alpha-Minimum Essential Medium (α-MEM, Gibco)
during osteogenic and adipogenic in in vitro differentiation of human supplemented with 10% heat inactivated FBS (Biowest) and 1% (v/v)
primary MSCs, determined the proteins and pathways most differently P/S (Invitrogen). All cells were expanded in a humidified atmosphere,
affected by miR-99a under pro-osteogenic conditions thought high- at 37 °C and 5% (v/v) CO2.

2
S.R. Moura, et al. Bone 134 (2020) 115303

2.4. Osteogenic differentiation Relative expression levels were calculated using the quantification
cycle (Cq) method, according to MIQE guidelines [33]. Data was ana-
To induce osteogenic differentiation, MC3T3 cells and hMSCs were lyzed using Bio-Rad CFX Manager software and all the reactions were
cultured in media supplemented with 10−7 M dexamethasone, 10−2 β- performed in duplicate.
glycerophosphate and 5 × 10−5 M ascorbic acid (all from Sigma-
Aldrich) and allowed to differentiate for 14 and 28 days, respectively. 2.8. miRNA mimics and inhibitor transfection
In parallel, cells were grown without osteogenic supplements (basal
conditions), as a control. RNA was collected at days 0, 3, 7 and 14 MC3T3 and RAW264.7 cells were transfected with 50 nM of miR-
(MC3T3) or 0, 3, 7, 14, 21 and 28 (hMSCs) of culture. Alkaline phos- 99a-5p mimics (mimics), miR-99a-5p inhibitor (inhibitor), or the re-
phatase (ALP) staining was performed at days 7 and 14 for MC3T3 and spective controls, namely miRNA mimics negative control (NC-mimics),
hMSCs, respectively; Alizarin Red S staining was performed at days 14 or miRNA inhibitor negative control (NC-inhibitor), all from Ambion,
and 28 for MC3T3 and hMSCs, respectively (Supplemental Methods). using Lipofectamine 2000 transfection reagent (Invitrogen), according
to manufacturer's instructions. Cells were incubated with the miRNA-
2.5. Adipogenic differentiation lipid complexes for 12 h and then collected for the distinct assays.
MC3T3-transfected cells were plated in osteogenic differentiation con-
To induce adipogenic differentiation, hMSCs from 6 donors were ditioned media, as described in 2.4, for ALP (day 7) and for Alizarin
cultured for 28 days with the following supplements: 10−4 M dex- (day 10) staining, RNA extraction (days 3 and 7), protein extraction
amethasone, 10−1 mM Indomethacin, 5 × 10−4 M IBMX and 10 μg/mL (day 7), collection of supernatants (day 7). MC3T3-transfected cells
insulin (all from Sigma-Aldrich). RNA was collected at days 0, 3, 7, 14, were also plated in growth condition media for evaluation of metabolic
21 and 28 of culture. Oil Red O staining was performed to confirm activity (Section 2.9), apoptosis and cell death (Section 2.10). RAW
adipogenic differentiation into mature adipocytes at day 21 and day 28 264.7-transfected cells were plated in osteoclastogenic differentiation
(Supplemental Methods). conditions, as described in Section 2.6, for TRAP staining (day 2) and
RNA extraction (day 2).
2.6. Osteoclastogenesis
2.9. Resazurin reduction assay
Osteoclastogenesis in human primary monocytes (6 donors) was
induced with 30 ng/mL human macrophage colony-stimulating factor To assess the effect of miR-99a-5p on cell metabolism/proliferation,
(M-CSF, ImmunoTools) and 50 ng/mL human receptor activator of MC3T3 cells transfected with either miR-99a-5p mimics, miR-99a-5p
nuclear factor kappa-Β ligand (RANKL, ImmunoTools). Cultures were inhibitor or the controls were seeded in 96-well plates. Resazurin redox
maintained during 21 days, and the media was carefully changed twice dye (0.01 mg/mL, Sigma-Aldrich) was added to the cells (10% (v/v)),
a week. RNA was collected at days 0, 7, 14 and 21. F-actin/DAPI and and incubated for 2 h at 37 °C. Experiments were performed for 7 re-
TRAP stainings were performed at days 7, 14 and 21 (Supplemental plicates in 5 time-points: days 0, 1, 2, 3 and 4. The supernatant was
Methods). RAW 264.7 cells were supplemented with 50 ng/mL mouse transferred to a dark wall 96 well plate and fluorescence levels fol-
RANKL (ImmunoTools). RNA was collected and TRAP staining assay lowing resazurin (non-fluorescent blue dye) reduction into resorufin
was performed each day, for 5 days. Multinucleated cells where con- (fluorescent and pink) were measured at an excitation wavelength of
sidered when ≥3 nuclei [29–32]. 530 nm and an emission wavelength of 590 nm, using the spectro-
photometer microplate reader Synergy MX (Biotek Synergy).
2.7. Reverse transcription and real-time quantitative polymerase chain
reaction (RT-qPCR) 2.10. Apoptosis assay by flow cytometry

Total RNA was extracted using TRIzol® Reagent (Invitrogen), ac- MC3T3 cells transfected with either miR-99a-5p mimics, miR-99a-
cording to the manufacturer's instructions. RNA concentration and 5p inhibitor or the respective controls were collected at day 7 post-
purity were determined using the NanoDrop Spectrophotometer ND- transfection. Cells were washed 3 times with PBS 1× and labelled with
1000 (ThermoFisher Scientific). RNA integrity was evaluated by gel FITC Annexin V Apoptosis Detection Kit I (BD Biosciences), according
electrophoresis. to manufacturer's instructions. Briefly, cells were re-suspended in
RNA was digested with TURBO DNA-free kit (Invitrogen), following binding buffer and incubated with FITC-Annexin V and Propidium
manufacturer's protocol, to remove potential DNA contaminants. Iodide (PI) for 15 min, at RT, in the dark. Then, the cells were run
Complementary DNA (cDNA) was synthetized using random hexamers within 1 h on the BD Accuri™ C6 system and the results were analyzed
(Invitrogen), dNTPs (Bioline) and SuperScript® III Reverse with FlowJo software.
Transcriptase kit (Invitrogen). qPCR reactions were performed using
cDNA, primers (Supplemental Table 2) and iQ SYBR Green Supermix 2.11. Enzyme-linked immunosorbent assay
(Bio-Rad) in a CFX Real-Time PCR Detection System (Bio-Rad) with the
following conditions: 3 min at 95 °C and 40 cycles of 30 s at 95 °C, 30 s The culture media from MC3T3 transfected cells was collected at
at 58 °C and 30 s at 72 °C. GAPDH was used as a reference gene. day 7 of osteogenic differentiation and centrifuged at 800g for 10 min,
miRNA expression was evaluated by RT-qPCR using TaqMan miRNA to remove cell debris. Osteoprotegerin (OPG) concentration in the su-
assays (Applied Biosystems). Briefly, cDNA was synthesized using total pernatants was determined using the commercial Quantikine ELISA kit
RNA, TaqMan MicroRNA Reverse Transcription Kit (Applied (MOP00 for OPG, R&D Systems), according to manufacturer's instruc-
Biosystems) and gene specific stem-loop Reverse Transcription primers tions. Absorbance was measured at 450 nm with correction set to
(hsa-miR-99a-5p and small nuclear RNA U6, Applied Biosystems), ac- 540 nm.
cording to the manufacturer's protocol. qPCR was carried out in CFX
Real-Time PCR Detection System (Bio-Rad), using cDNA, 2.12. Effect of osteogenic-conditioned media on osteoclastogenesis
SsoAdvanced™ Universal Probes Supermix (Bio-Rad) and hsa-miR-99a-
5p or small nuclear RNA U6 TaqMan assays (Applied Biosystems), Cell culture media derived from MC3T3 cells transfected with miR-
under the following conditions: 10 min at 95 °C, 40 cycles of 15 s at 99a-5p mimics, miR-99a-5p inhibitor, or the respective controls, was
95 °C and 1 min at 60 °C. Small nuclear RNA U6 was used as a reference collected following 7 days of culture in osteogenic-inducing conditions.
gene. Media was centrifuged at 800g, for 10 min, at 4 °C, to remove cells and

3
S.R. Moura, et al. Bone 134 (2020) 115303

cell debris. RAW 264.7 cells were incubated with the MC3T3-derived 2.14. Western blot
cell culture media at a proportion of 3:1 (3 volumes of conditioned
media and 1 volume of α-MEM with FBS) for 72 h. As a control, RAW Protein extracts (30 μg of protein) were prepared in reducing
264.7 cells were also independently incubated with α-MEM + 10% (v/ loading buffer, denatured at 95 °C for 5 min, and resolved in 10% (v/v)
v) FBS (negative control) or α-MEM + 10% (v/v) FBS + 50 ng/mL polyacrylamide SDS-PAGE gels at 100 V, along with the molecular
RANKL (positive control). Expression of osteoclastogenic associated weight marker. Proteins were wet-transferred to nitrocellulose mem-
genes was evaluated by RT-qPCR and the number of multinucleated branes, blocked with a 5 g/L non-fat dry milk solution, and probed
cells was quantified by microscopy. overnight at 4 °C using the following primary antibodies: anti-
Fibromodulin (H-11, Santa Cruz Biotechnology, 1:500), anti-Lumican
(B-9, Santa Cruz Biotechnology, 1:500), and anti-GAPDH (Cell
2.13. Proteomic analysis Signalling, 1:1000). Next, the membranes were probed with an ap-
propriate secondary antibody conjugated to horseradish peroxidase
MC3T3 cells transfected with miR-99a-5p mimics, miR-99a-5p in- (HRP, GE Healthcare, 1:10000) for 1 h, at RT. After washing, mem-
hibitor or the respective controls were cultured in osteogenic differ- branes were incubated with chemiluminescent substrate (ECL, GE
entiation conditions. Cells were harvested at day 7 of differentiation Healthcare), for signal detection in autoradiographic films (GE
and washed 6 times with cold PBS 1×, before being lysed with cold Healthcare). Protein bands were quantified on Fiji, and relative protein
RIPA buffer [20 mM Tris-HCl pH 7.5, 150 mM NaCl, 1% (v/v) Triton X- levels were calculated using GAPDH bands as normalizer.
100 and 1% (v/v) NonidetP-40] in the presence of protease inhibitors
[(10 μL/mL phenylmethylsulfonyl fluoride (PMSF, 200 mM), 10 μL/mL 2.15. Statistical analysis
Leupeptin (1 ng/mL) and 10 μL/mL Aprotinin (1 mg/mL)] and phos-
phatase inhibitors [20 μL/mL NaVO3 (50 nM), 50 μL/mL Na4P2O7 Statistical data analysis was performed on Prism 7 (GraphPad
(50 mg/mL) and 10 μL/mL NaF (10 mM)], for 30 min. Cell lysates were Software, Inc.). Gaussian distribution was tested by the Shapiro-Wilk
centrifugation at 20,000g, for 10 min, at 4 °C and the supernatant- and Kolmogorov-Smirnov normality tests. Data was only considered to
containing protein was quantified using DC protein assay kit (Bio-Rad). follow a normal distribution when passed both tests (alpha = 0.05). For
Protein identification and label-free quantitation was performed by non-normal distribution data, or when n < 5, non-parametric tests
nanoLC-MS/MS, composed by an Ultimate 3000 liquid chromatography were used to evaluate significant differences between samples, namely
system coupled to a Q-Exactive Hybrid Quadrupole-Orbitrap mass two-tailed Wilcoxon matched pairs test (between 2 groups) or Friedman
spectrometer (Thermo Scientific, Bremen, Germany). Samples were test (> 2 groups) followed by uncorrected Dunn's multiple comparison
loaded onto a trapping cartridge for 3 min and further separated on an test. When the data passed normality tests, one-way ANOVA (> 2
nano-C18 column at 300 nL/min. Peptide separation gradient was the groups), followed by Sidak's multiple comparison or Turkey's multiple
following (A: 0.1% (v/v) FA, B: 80% (v/v) ACN 0.1% (v/v)): 5 min comparisons tests were used. For the RT-qPCR data, the fold change
(2.5% (v/v) B to 10% (v/v) B), 100 min (10% (v/v) B to 35% (v/v) B), differences (2−∆∆CT) were tested using the Wilcoxon signed-ranked test.
20 min (35% (v/v) B to 55% (v/v) B), 3 min (55% (v/v) B to 99% (v/v) Statistical significance was considered for P < 0.05 (* P < 0.05; **
B) and 12 min (hold 99% (v/v) B). Data acquisition was controlled by P < 0.01 and *** P < 0.001).
Xcalibur and Tune software (Thermo Scientific).
The mass spectrometer was operated in data-dependent positive 3. Results
acquisition mode alternating between a full scan (m/z 380–1580) and
subsequent HCD MS/MS of the 10 most intense peaks from full scan. 3.1. miR-99a-5p is downregulated during early stages of osteogenic
Raw data was processed using Proteome Discoverer 2.3.0.523 software differentiation of human primary MSCs and in mouse MC3T3 cell line
(Thermo Scientific). Protein identification was performed with Sequest
HT search engine against the Mus musculus entries from the UniProt When hMSCs were induced to differentiate into the osteogenic
database (https://www.uniprot.org/). Mass tolerance was 10 ppm for lineage, the ALP staining and the mineralized area, assessed by Alizarin
precursor and 0.02 Da for fragment ions, respectively. Maximum al- staining, were significantly increased when compared to basal condi-
lowed missing cleavage sites was set 2. Cysteine carbamidomethylation tions (Fig. 1A and B). The expression profile of miR-99a-5p in 6 hMSCs
was defined as constant modification. Methionine oxidation and protein donors shows that miR-99a-5p levels were significantly downregulated
N-terminus acetylation were defined as variable modifications. Protein at days 3, 7 and 14 of osteogenic differentiation compared to the basal
and peptide confidence were set to high. The processing node control (P < 0.05, days 3, 7 and 14), while at days 21 and 28 miR-99a-
Percolator was enabled with the following settings: maximum delta Cn 5p expression was restored to basal levels (Fig. 1C). This indicates a
0.05; decoy database search target FDR 1%, validation of based on q- potential effect of this miRNA in the early stages of hMSCs osteogenic
value. Analyzed samples were normalized against to the total peptide differentiation rather than in the later mineralization stages. In the
signal in each experiment and its quantitative evaluation was achieved MC3T3 mouse pre-osteoblastic cell line, osteogenic differentiation also
by pairwise comparisons of the detected peptides. Data was corrected significantly increased ALP staining, and the deposition of calcium at
using Benjamin Hochberg method. Protein levels were compared by days 7 and 14, respectively (Fig. 1D and E). RT-qPCR results show that
using the median ratio. miR-99a-5p was significantly downregulated during osteogenic differ-
Bioinformatics and data analysis was performed for proteins ex- entiation, specifically at day 7 (P < 0.05), which is in agreement with
pressed in the two independent experiments with fold-change (FC) our previously published microarray results [11], and at day 14
differences ≥1.25 or ≤−1.25. A minimum of two unique peptides (P < 0.05), compared to the basal control in each time point (Fig. 1F).
were required for further evaluation. The differentially expressed pro-
teins were classified according to Gene Ontology (GO) annotations. The 3.2. miR-99a-5p is upregulated during the early stages of adipogenic
enriched Biological Process analysis was performed using Protein differentiation of human primary MSCs
ANalysis THrough Evolutionary Relationships (PANTHER) tool.
miRWalk2.0 database was used for miRNA target prediction (http:// To determine the impact of miR-99a-5p in the hMSCs' adipogenesis,
mirwalk.umm.uni-heidelberg.de/). For the analysis of the top 20 pro- we induced their commitment into the adipogenic lineage. Results show
teins with opposite profiles between the miR-99a-mimics/NC-mimics an increased Red O staining of the lipid droplets in terminally differ-
and in the miR-99a-5p-inhibitor/NC-inhibitor groups, only the proteins entiated mature adipocytes, after 21 and 28 days of culture in adipo-
with greater than or equal to 6 unique peptides were considered. genic-inducing conditions, compared to basal controls (Fig. 2A). The

4
S.R. Moura, et al. Bone 134 (2020) 115303

Fig. 1. miR-99a-5p expression during osteogenic differentiation of hMSCs and MC3T3 cells. A) Representative image of ALP staining at day 14 and detection of
calcium deposits by Alizarin Red S staining (stained red) at day 28 in hMSCs under osteogenic differentiation (Dif) and basal conditions (5×, scale 500 μm). B) ALP
staining and Alizarin quantification in hMSCs after 14 and 28 days, respectively (median, N = 6, two-tailed Wilcoxon matched pairs test, *P < 0.05). C) miR-99a-5p
expression in hMSCs after 3, 7, 14, 21 and 28 days (tdif) of culture in osteogenic versus basal conditions (median, N = 6, Wilcoxon signed ranked test, * P < 0.05).
D) Representative image of ALP staining at day 7 and detection of calcium deposits by Alizarin Red S staining (stained red) at day 14 in MC3T3 cells under osteogenic
differentiation and basal conditions (5×, scale 500 μm). E) ALP staining and Alizarin Red S quantification in MC3T3 after 7 and 14 days, respectively (median,
N = 6, two-tailed Wilcoxon matched pairs test, * P < 0.05). F) miR-99a-5p expression in MC3T3 pre-osteoblasts after 3, 7 and 14 days (tdif) of culture in osteogenic
versus basal condition (median, N = 6, Wilcoxon signed ranked test, * P < 0.05). (For interpretation of the references to colour in this figure legend, the reader is
referred to the web version of this article.)

expression of the adipogenic-specific markers ADIPOQ (Adiponectin) negatively with osteogenic differentiation.
and PPARG2 (Peroxisome proliferator activated receptor gamma) was
upregulated after 21 and 28 days (P < 0.05) (Fig. 2B). Interestingly, at
3.3. miR-99a-5p regulates negatively the osteogenic differentiation but has
early time points of adipogenic differentiation, miR-99a-5p was sig-
no effect on proliferation or apoptosis of MC3T3 cells
nificantly upregulated (P < 0.05, days 3 and 7), compared with basal
conditions (Fig. 2C), while decreased at days 21 and 28 (P < 0.05).
Next, we evaluated the biological impact of miR-99a-5p through
Therefore, at early stages of hMSCs differentiation, miR-99a-5p ex-
gain- and loss-of-function studies in MC3T3 cells, which were efficiently
pression levels associate positively with adipogenic differentiation, but
transfected (Fig. 3A). miR-99a-5p overexpression decreased ALP

5
S.R. Moura, et al. Bone 134 (2020) 115303

Fig. 2. miR-99a-5p expression during adipogenic differentiation of hMSCs.


A) Representative image of Oil Red O staining (lipids stained red) at days 21 and 28 under adipogenic differentiation and basal conditions (10×, scale 200 μm; 5×,
scale 500 μm, respectively). B) Expression of adipogenic specific genes, ADIPOQ and PPARG2 (median, N = 6, Wilcoxon signed ranked test, * P < 0.05). C) miR-
99a-5p expression in hMSCs grown in adipogenic-inducing conditions relative to basal conditions after 3, 7, 14, 21 and 28 days (tdif) in culture (median, N = 6,
Wilcoxon signed ranked test,* P < 0.05). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

staining in 42.20% (P < 0.01), whereas miR-99a-5p inhibition in- overexpressing cells cultured under osteogenic inducing conditions at
creased ALP staining in 70.77% (P < 0.05), after 7 days in osteogenic days 3 and 7 (Fig. 3D). Conversely, Alp and Runx2 were significantly
inducing conditions, compared to controls (Fig. 3B). Also, results show increased at day 3 (P < 0.05) and day 7 (P < 0.05) in response to
a 60.29% decrease (P < 0.001) on the formation of calcium deposits in miR-99a-5p inhibition (Fig. 3D). In order to understand if the effect of
miR-99a-mimics transfected cells, and a 88.64% increase (P < 0.01) in miR-99a-5p on osteogenic differentiation could be a consequence of its
miR-99a-inhibitor transfected cells, both after 10 days of differentiation impact on cell proliferation, the metabolic activity was measured as an
and compared to their respective controls (Fig. 3C). In agreement with indirect method to evaluate cell proliferation. Results show that neither
these results, mRNA levels of key the osteogenic markers Alpl and upregulation nor downregulation of miR-99a-5p had a significantly
Runx2 were significantly decreased (P < 0.05) in miR-99a-5p effect on cell proliferation compared to controls (Fig. 3E). The Annexin

6
S.R. Moura, et al. Bone 134 (2020) 115303

(caption on next page)

7
S.R. Moura, et al. Bone 134 (2020) 115303

Fig. 3. Effect of miR-99a-5p in osteogenic differentiation, cell proliferation and apoptosis. A) miR-99a-5p expression after transfection of MC3T3 cells with miR-99a
mimics, miR-99a inhibitor or respective controls (miR-99a NC-mimics, miR-99a NC-inhibitor), and culture in osteogenic-inducing conditions (median, N = 7,
Wilcoxon signed ranked test, * P < 0.05). B) ALP and C) Alizarin staining after transfection and culture for 7 and 10 days, respectively, in osteogenic-inducing
conditions (mean ± SD, N = 7, two-tailed Wilcoxon matched pairs test, * P < 0.05, ** P < 0.01 and *** P < 0.001). D) Alpl and Runx2 expression levels at days
3 and 7 after transfection and incubation in osteogenic-inducing conditions (median, N = 7, Wilcoxon signed ranked test, * P < 0.05). E) Representative fluor-
escence profile (resorufin) of 4 independent experiments measured every 24 h for 4 days in transfected cells (N = 4, Friedman test, followed by uncorrected Dunn's
multiple comparison test; n.s.: non-significant). F) Representative image of flow cytometry analysis of AnnexinV/PI staining and quantification of viable (Annexin
V− PI−), early apoptotic (Annexin V+ PI−), or late apoptosis and dead cells (Annexin V− PI+ and Annexin V+ PI+) (median, N = 3, Friedman test, followed by
uncorrected Dunn's multiple comparison test, n.s.: non-significant).

V/PI staining was used to evaluate apoptosis at days 3 and 7. Flow Globally, this high-throuput proteomic analysis in MC3T3-trans-
cytometry results show that overexpression or inhibition of miR-99a-5p fected cells shows that distinct intracellular proteins related to osteo-
in MC3T3 cells did not significantly affect early apoptosis or late genic differentiation, as well as matrix proteins, are modulated by miR-
apoptosis/cell death (Fig. 3F). Taken together, the results indicate that 99a-5p.
miR-99a-5p exerted an osteogenic-suppressive effect in MC3T3 cells in
vitro, but has no effect on proliferation or apoptosis.
3.5. miR-99a-5p regulates osteoclastogenesis through a paracrine effect
3.4. Modulation of miR-99a-5p levels changes cellular proteomic profile
Bone homeostasis requires the crosstalk between osteoblasts and
To identify novel miR-99a-5p targets and elucidate relevant me- osteoclasts, which involves secreted mediators. As miR-99a-5p may
chanisms and/or pathways regulated by this miRNA, cells transfected affect levels of secreted protein, we analyzed the expression levels of
with miR-99a-5p-mimics, miR-99a-5p-inhibitor, or the respective con- the two main mediators of osteoblast to osteoclast communication:
trols, were analyzed by mass spectrometry-based proteomics. OPG, as a protector of excessive bone resorption; RANKL, as a positive
Proteomic profiling identified and quantified a total of 3907 proteins. regulator of bone resorption [2]. qRT-PCR results show that Tnfrsf11b
Analysis revealed differences in 1736 proteins when comparing miR- mRNA (OPG encoding gene) is decreased by miR-99a-5p over-
99a-5p overexpressing cells versus NC-mimics transfected cells expression (P < 0.05), while increased by miR-99a-5p inhibition
(FC ≥ |1.25|), and in 809 proteins when comparing miR-99a-5p in- (P < 0.05, Fig. 5A). On the other hand, Tnfrsf11 (RANKL encoding
hibited cells versus NC-inhibitor transfected cells (FC ≥ |1.25|). Out of gene) expression does not change following miR-99a-5p mimics/in-
those, 64 differentially expressed proteins were simultaneously down- hibitor transfection (Fig. 5A). Interestingly, in the control conditions,
regulated by miR-99a-5p mimics and upregulated by miR-99a-5p in- Tnfrsf11b is expressed at much higher levels than Tnfrsf11. Therefore,
hibitor. When only considering the statistically different proteins Tnfrsf11b is the main contributor to the decreased in Tnfrsf11b/Tnfrsf1
(P < .05) that were downregulated by the miR-99a-5p mimics or expression ratio by the miR-99a-5p-mimics, and the increase by miR-
upregulated by miR-99a-5p inhibitor, results show 44 significantly 99a-5p-inhibitor (P < 0.05). Next, we investigated whether the OPG
different proteins (Supplemental Table 3). In silico analysis (miRWalk protein levels in the conditioned media were altered. In agreement with
2.0. database [34]) revealed that 6 of those proteins were predicted by the gene expression levels, results show that secreted OPG levels are
at least one algorithm to originate from transcripts with putative miR- significantly reduced in supernatants of miR-99a-5p-overexpressing
99a-5p binding sites (Supplemental Table 4). Conversely, 316 proteins cells (P < 0.001), while increased in supernatants of miR-99a-5p-in-
were upregulated in miR-99a-overexpressing cells and downregulated hibitor transfected cells (P < 0.001) (Fig. 5B).
in miR-99a-5p-knowckdown cells (Fig. 4A). Next, we tested the effect of MC3T3-conditioned media on osteo-
Gene ontology analysis shows that the molecular function mostly clast differentiation, following miR-99a-5p overexpression or inhibi-
affected by changes in miR-99a-5p expression, either by its over- tion. RAW 264.7 cells were cultured during 2 days with media derived
expression or inhibition, is the cellular catalytic activity (GO: 0003824, from MC3T3-transfected with miR-99a-5p-mimics or with miR-99a-5p-
Fig. 4B). Included in this term, the protein kinase activity (GO: inhibitor, in the absence of RANKL supplementation. As a positive
0004672) shows the highest percentage of gene hits compared to the control, RAW 264.7 cells were independently treated with RANKL.
total number of function hits (Supplemental Table 5). Fig. 4C shows the Remarkably, the number of multinucleated cells showed an increase by
20 most differently expressed proteins with opposite regulation by miR- 33.6% (P < 0.05) and a decrease by 32.8% (P < 0.001), following
99a-5p mimics and its inhibitor. These include extracellular matrix culture with the conditioned media derived from miR-99a-5p-mimics or
(ECM) components that are members of the small interstitial leucine- miR-99a-5p-inhibitor transfected MC3T3 cells, respectively (Fig. 5C).
rich repeat proteoglycans family (SLRPs), namely Fibromodulin (Fmod) The expression levels of osteoclast fusion markers Dcstamp, Ccl2, and
and Lumican (Lum), which are increased in miR-99a-5p-inhibitor osteoclastogenic-specific marker Ctsk, were increased in RAW 264.7
transfected cells (FC = 1.90 and 1.25, respectively) and decreased by cells stimulated by miR-99a-5p-mimics-MC3T3 conditioned media
miR-99a-5p-mimics transfected cells (FC = −1.44 and −1.38, re- (Dcstamp: FC = 1.56, P < 0.05; Ccl2: FC = 1.85, P < 0.05; Ctsk:
spectively) (Fig. 4C). This effect was validated by western blot for Fmod FC = 1.57, P < 0.05), whereas the opposite effect was observed when
(Fig. 4D), but not for Lum (Supplemental Fig. 2), which is in agreement cells were exposed to miR-99a-5p-inhibitor-MC3T3 conditioned media
with the higher differential effect found for Fmod by mass spectometry (Dcstamp: FC = 0.745, P < 0.05; Ccl2: FC = 0.862, P < 0.05; Ctsk:
analysis (Fig. 4C). Additionally, Fmod mRNA expression was reduced by FC = 0.835, P < 0.05) (Fig. 5D). These results suggest that the con-
miR-99a-5p overexpression (P < 0.05), while increased by miR-99a- ditioned media from miR-99a-5p transfected cells influences the dif-
5p inhibition (P < 0.05) (Fig. 4D). ferentiation of osteoclast-like cells by affecting the expression of os-
Most of the differently expressed proteins were detected after miR- teoclastogenic markers, as well as their ability to form multinucleated
99a-5p overexpression. According to GO term analysis, 76 of those cells.
proteins are associated with “cell differentiation” and, of those, the top Interestingly, intracellular levels of miR-99a-5p in RAW 264.7 cells
20 include the osteogenic-specific markers ALPL (FC = −2.10), were significantly increased after culture with miR-99a-5p-MC3T3
RUNX2 (FC = −2.07), and Osterix (SP7, FC = −1.53). Additionally, conditioned media and decreased after culture with anti-miR-99a-5p-
changes were also found in proteins with roles both in “cell differ- MC3T3 conditioned media (Supplemental Fig. 3). Taken together these
entiation” and in “extracellular matrix production”, COL8A1 results support a role for miR-99a-5p in osteoclastogenesis via a para-
(FC = −1.82) and COL4A1 (FC = −1.60) (Fig. 4E). crine effect.

8
S.R. Moura, et al. Bone 134 (2020) 115303

(caption on next page)


9
S.R. Moura, et al. Bone 134 (2020) 115303

Fig. 4. Proteomic analysis of miR-99a-5p downstream targets and processes. A) Number of differently expressed proteins in MC3T3 cells after transfection with miR-
99a-5p mimics, inhibitor or respective controls (FC ≥ |1.25|). Number of proteins significantly increased by miR-99a inhibition or decreased by miR-99a over-
expression (FC ≥ |1.25| and P < 0.05) with miR-99a-5p predicted binding sites by miRWalk 2.0. B) Gene ontology enrichment analysis, according to PANTHER
classification system. The Y-axis shows the number of proteins associated with determined GO term (X-axis). C) Levels (fold change) of the most antagonistic proteins
between miR-99a-5p-inibition and miR-99a-5p-overexpressing cells (≥6 unique peptides). D) Fibromodulin (FMOD) protein levels by Western blot in MC3T3 cells
after transfection with miR-99a-5p mimics, inhibitor, or respective controls, and following 7 days of culture in osteogenic-inducing conditions. GAPDH was used as a
normalyzer. Representative image (left panel) and quantification FMOD levels of 5 independent experiments using FIJI software (quantification FMOD/GAPDH
expression in 5 independent experiments (mean ± SD, N = 5, one-way ANOVA, followed by Sidak's multiple comparison test, * P < 0.05). Fmod mRNA expression
levels (right panel, median, N = 6, Wilcoxon signed ranked test, * P < 0.05). E) List of the top 20 most differential proteins associated with the GO term “cell
differentiation” and decreased by miR-99a-5p-overexpressing cells.

Fig. 5. Effect of conditioned media from miR-99a-5p mimics/inhibitor MC3T3-transfected cells on osteoclastogenesis. A) Expression of Tnfrsf11b, Tnfsf11 and
Tnfrsf11b/Tnfsf11 ratio in MC3T3-transfected cells with either miR-99a-5p mimics, inhibitor or respective controls, at day 7 of osteogenic differentiation (median,
N = 6, Wilcoxon signed ranked test, * P < 0.05, n.s.: non-significant). B) OPG protein levels in conditioned media from MC3T3-transfected cells relative to control
(mean ± SD, N = 8, one-way ANOVA, followed by Sidak's multiple comparison test, *P < 0.05). C) Number of multinucleated cells (mean ± SD, N = 6, one-way
ANOVA, followed by Sidak's multiple comparison test, * P < 0.05, ** P < 0.01 and *** P < 0.001) and E) expression of osteoclastogenic-markers Dcstamp, Ccl2
and Ctsk (median, N = 6, Wilcoxon signed ranked test, * P < 0.05) of RAW 264.7 cells grown in the conditioned media from MC3T3-transfected cells.

10
S.R. Moura, et al. Bone 134 (2020) 115303

(caption on next page)

11
S.R. Moura, et al. Bone 134 (2020) 115303

Fig. 6. Time-dependent increase of miR-99a-5p during monocyte to osteoclast differentiation. A) Osteoclastogenesis of human monocytes. F-actin and nuclei staining
at days 7, 14 and 21 of differentiation (supplemented with MCS-F and RANKL) and in control cells (MCS-F) (10×, scale 100 μm). B) Number of cells with 1, 2, 3–9
and > 10 nuclei per cell across 6 separate random fields (median, N = 6, Friedman test, followed by uncorrected Dunn's multiple comparison test, * P < 0.05 and **
P < 0.01). C) Average of the area of the human osteoclasts after 7, 14 and 21 days of differentiation (mean + SD, N = 6, one-way ANOVA, followed by Turkey's
multiple comparison test, ** P < 0.01 and *** P < 0.001). D) Expression levels of human osteoclast specific genes ACP5, CTSK and CALCR in the presence of
osteoclastogenic-inducing supplements (median, N = 6, Wilcoxon signed ranked test, * P < 0.05). E) miR-99a-5p expression levels in human monocyte-derived
osteoclasts grown in osteoclastogenic-inducing conditions after 7, 14 and 21 days, relative to day 0/non-stimulated monocytes (median, N = 6, Wilcoxon signed
ranked test, * P < 0.05). F) miR-99a-5p expression levels of RAW 264.7 cells grown in osteoclastogenic-inducing conditions after 1, 2, 3, 4 and 5 days, relative to
day 0 (median, N = 6, Wilcoxon signed ranked test, * P < 0.05).

3.6. miR-99a-5p expression increases during osteoclastogenesis of human 4. Discussion


primary monocytes and of mouse RAW 264.7 cells
Cell lineage commitment is crucial when considering the potential
The results obtained this far prompted us to investigate the levels of use of miRNA-based engineered MSCs for bone diseases or fractures [6].
miR-99a-5p during osteoclastogenesis. As such, human primary Herein, we show that, in human primary bone-marrow-derived MSCs,
monocytes isolated from healthy blood donors were seeded in the miR-99a-5p expression is decreased during early stages of osteogenic
presence of RANKL and M-CSF, and the number of multinucleated cells differentiation, and that miRNA inhibition promotes a pro-osteogenic
was quantified over time. Monocytes stimulated with only M-CSF were profile. The miR-99a-5p expression profile during osteogenic differ-
used as a negative control (Fig. 6A, Supplemental Fig. 4). As expected, entiation is in agreement with our published microarray data in a
the number of nuclei per cell increases in a time-dependent manner mouse pre-osteoblastic cell line [11] and with Tang et al. findings in
(Fig. 6B). While the number of cells with only one or two nuclei pro- mice MSCs [22]. Furthermore, we show, for the first time in human
gressively decreases during differentiation into OCs at day 21 (1 nuclei: MSCs, that miR-99a-5p increases in early stages of adipogenic differ-
P < 0.01; 2 nuclei: P < 0.05), the number of cells with > 10 nuclei is entiation, suggesting an opposite expression profile in comparison with
significantly higher at day 21 (P < 0.01), compared to day 7 of dif- osteogenic differentiation. Interestingly, in distinct cancer cell lines,
ferentiation (Fig. 6B). Accordingly, along differentiation the cells were miR-99a-5p has been shown to act as a tumour suppressor through
able to merge and originate large OCs, increasing the cells size over negative regulation of proliferation (e.g. Bladder [35], esophageal
time (day 14: P < 0.001; day 21: P < 0.01) (Fig. 6C). Also, the ex- squamous cell [36], and renal cell carcinoma [37]) and inducing
pression levels of the osteoclastogenic markers ACP5 (TRAP encoding apoptosis (e.g. breast cancer [38]). However, in this study, we show
gene), CTSK, and CALCR (CTR encoding gene) were significantly in- that in a non-tumoral cell line (MC3T3), miR-99a-5p does not affect
creased over time (Fig. 6D), confirming a successful osteoclast differ- proliferation or apoptosis/cell death. Therefore, the miR-99a-5p-medi-
entiation for all 6 donors. Finally, the expression of miR-99a-5p was ated control of osteogenic differentiation is a specific effect and not an
evaluated during osteoclastogenesis in these cells. As illustrated in indirect consequence of cell proliferation/apoptosis. This study shows
Fig. 6E, miR-99a-5p is significantly overexpressed during the differ- that levels of miR-99a-5p are timely controlled during human in vitro
entiation of human monocyte-derived OCs (day 7: FC = 4.12, osteogenic/adipogenic differentiation and during osteoclastogenesis.
P < 0.05; day 14: FC = 6.31, P < 0.05; day 21: FC = 10.52, However, it does not uncover the upstream mechanisms that regulate
P < 0.05), compared to unstimulated cells (day 0). Similar to the the expression of this miRNA, which can occur at multiple levels.
human monocytes-derived OCs, miR-99a-5p is upregulated during os- Generally, transcriptional and post-transcriptional modifications can
teoclastogenesis of RAW 264.7 murine monocytes, particularly at the control miRNA expression [39]. Transcriptional control can be depen-
later time points (day 4: FC = 5.20, P < 0.05; day 5: FC = 6.35, dent or independent of the miRNA host gene, while post-transcriptional
P < 0.01; (Fig. 6F). The expression of osteoclastogenic genes was also modifications are dependent of the miRNA processing and stability.
evaluated during 5 days of RANKL-induced differentiation in RAW Additionally, the expression of a given miRNA can be influenced by cell
264.7 cells (Supplemental Fig. 5). endogenous factors, like their production of cytokines, and exogenous
stimuli that cells may be exposed to, such as chemical compounds [39].
The assessment of miRNA expression during osteogenic/adipogenic
3.7. miR-99a-5p positively regulates osteoclastogenesis differentiation and during osteoclastogenesis described here, was per-
formed in human primary cells, while miRNA gain- and loss-of-function
To investigate the role of miR-99a-5p in osteoclast differentiation, experiments were performed using mouse cell lines. Nonviral gene
gain- and loss-of-functions studies were performed in RAW 264.7 cells. delivery (transfection/electroporation) in primary cells (either MSCs or
Levels of miR-99a-5p were increased upon transfection with miR-99a- monocytes) presents low efficiency and toxic effects, while viral
5p mimics and decreased after transfection with its inhibitor (Fig. 7A). mediated delivery presents safety issues and unexpected side effects,
Two days after transfection, and in the presence of RANKL stimulation, which is particularly relevant when envisioning a future in vivo ap-
miR-99a-5p-overexpressing RAW 264.7 cells show a significant increase plication [40]. This is a limitation of this study, as we cannot exclude
in Dcstamp expression (P < 0.05), and a decrease with the inhibition of that the observed effects could be cell line or species specific. Still, miR-
miR-99a-5p (P < 0.05) (Fig. 7B). Also, Ccl2 is upregulated by miR-99a 99a-5p is evolutionary conserved between human and mouse (http://
(P < 0.05). Although not statistically significant, expression of Ctsk www.mirbase.org/), and it mature sequence shows 100% homology
follow the same tendency, with an upregulation of by miR-99a-mimics between these species. RAW 264.7 cells have been widely used to study
(P = 0.125) and a reduction by miR-99a-inhibitor (P = 0.094) monocyte/osteoclast differentiation in vitro because these cells are easy
(Fig. 7B). Finally, we tested the ability of these cells to fuse and form to culture, to genetically manipulate, and can overcome the limited
osteoclast-like cells under RANKL stimulation. Results show that miR- lifespan of primary cells. RAW264.7 present similarities with the pri-
99a-5p overexpression promotes the formation of multinucleated cells mary monocytes, for instance, both require RANKL supplementation for
(FC = 1.54, P < 0.05), while miR-99a-5p inhibition impairs cell fu- in vitro differentiation. Nevertheless, their usage should be carefully
sion (FC = 0.776, P < 0.05) (Fig. 7C). Taken together, these results considered. High-throughput proteomics analysis performed by Ng
demonstrate that miR-99a-5p has a functional effect on osteoclasto- et al. shows a distinct protein profile during osteoclast formation be-
genesis. tween RAW 264.7 and mice bone marrow macrophages [41]. The most

12
S.R. Moura, et al. Bone 134 (2020) 115303

Fig. 7. miR-99a-5p promotes osteoclastogenesis of RAW 264.7 cells.


A) miR-99a-5p expression in RAW 264.7-transfected cells with miR-99a mimics, miR-99a inhibitor, and respective controls (median, N = 6, Wilcoxon signed ranked
test, * P < 0.05). B) Dcstamp, Ctsk and Ccl2 expression levels 3 days after transfection and culture in the presence of RANKL (median, N = 6, Wilcoxon signed ranked
test, * P < 0.05). C) Number of multinucleated cells (≥3 nuclei) 3 days after transfection and culture in the presence of RANKL, relative to controls (mean ± SD,
N = 6, one-way ANOVA, followed by Sidak's multiple comparison test, * P < 0.05 and ** P < 0.01).

differently affected biological processes are cell cycle, cytoskeleton re- target a single mRNA. Therefore, our approach has 1) resulted in a
organization and apoptosis [41]. Additionally, RAW 264.7 cells do not global analysis on the molecular and biological processes most affected
exclusively differentiate into osteoclast but also in other multinucleated by miR-99a-5p expression, 2) identified novel players regulated by miR-
cells, including macrophage polykaryons [42]. To mitigate some of 99a-5p during osteogenic differentiation, as well as 3) detected the
these limitations, our study not only evaluated the number of multi- impact of miR-99a-5p on known key osteogenic markers. However, this
nucleated cells but also tested the levels of key osteoclastogenic mar- analysis was performed in proteins isolated from cell lysates and thus,
kers. excludes secreted proteins present in the supernatant. To partially
An advantage of miRNA-based therapies is that a single miRNA can overcome this limitation, and considering the importance of RANK/
affect multiple targets and simultaneously control distinct pathways. In RANLK/OPG pathway in bone remodeling, the secreted levels of OPG
fact, our proteomic analysis shows that several cell differentiation and were measured. OPG, an endogenous anti-osteoclastogenic decoy for
extracellular matrix classes of proteins are affected by miR-99a-5p le- RANKL [45,46], was increased by miR-99a-5p inhibition, while abro-
vels. Among these, a class II small leucine-rich proteoglycan was gated by miR-99a-5p mimics. As expected, OPG could only be detected
identified to be regulated by this miRNA, namely Fmod. In the litera- by ELISA and was not identified in the cell proteomic analysis [47].
ture, Fmod is described as a promotor of mineralization [43] and its Importantly, OPG is a central player in bone metabolism and potentially
ablation negatively affects bone mineral density [44]. The advantage of a therapeutic candidate. In vivo studies showed that OPG is able to
using a high-throughput screening method is that we can identify many reverse osteoporosis and that OPG-engineered MSCs are able to pro-
distinct proteins and pathways associated to the expression of a specific moted critical-size bone repair in an osteoporosis mice model [48,49].
miRNA. This is particularly relevant when studying miRNA, since these Nonetheless, our analysis does not excluded the involvement of other
small ncRNAs are known to have multiple targets, contrary to silencing mediators that may be present in the supernatant of MC3T3-transfected
RNAs (siRNA) that are exogenous sequences specifically designed to cells that could be able to impact the communication with osteoclast.

13
S.R. Moura, et al. Bone 134 (2020) 115303

These may include different mediators like cytokines/chemokines, or remodeling, Arch. Biochem. Biophys. 473 (2) (2008) 139–146, https://doi.org/10.
even nucleic acids, such as miRNAs encapsulated or not in extracellular 1016/j.abb.2008.03.018.
[3] A.E. Kearns, S. Khosla, P.J. Kostenuik, Receptor activator of nuclear factor kappaB
vesicles [50–52]. ligand and osteoprotegerin regulation of bone remodeling in health and disease,
Therefore, the effect that the total secretome produced by miR-99a- Endocr. Rev. 29 (2) (2008) 155–192, https://doi.org/10.1210/er.2007-0014.
5p/anti-miR-99a-5p-transfected MC3T3 cells have on osteoclastogen- [4] J.A. Kanis, E.V. McCloskey, H. Johansson, A. Oden, Approaches to the targeting of
treatment for osteoporosis, Nat. Rev. Rheumatol. 5 (8) (2009) 425–431, https://doi.
esis was investigated. Surprisingly, supernatants derived from anti-miR- org/10.1038/nrrheum.2009.139.
99a-5p cells could not only reduce osteoclastogenesis but also decrease [5] X. Feng, J.M. McDonald, Disorders of bone remodeling, Annu. Rev. Pathol. 6 (2011)
the levels of miR-99a-5p expression by the osteoclasts, suggesting a role 121–145, https://doi.org/10.1146/annurev-pathol-011110-130203.
[6] J. Freitas, S.G. Santos, R.M. Goncalves, J.H. Teixeira, M.A. Barbosa, M.I. Almeida,
for this miRNA in osteoclasts. Herein, we show that Genetically engineered-MSC therapies for non-unions, delayed unions and critical-
miR-99a-5p is increased during osteoclastogenesis of primary size bone defects, Int. J. Mol. Sci. 20 (14) (2019), https://doi.org/10.3390/
human monocytes and that miR-99a-5p promotes the differentiation of ijms20143430.
[7] Q. Chen, P. Shou, C. Zheng, M. Jiang, G. Cao, Q. Yang, J. Cao, N. Xie, T. Velletri,
osteoclasts. This is in line with De la Rica et al. [53] and Franceschetti
X. Zhang, C. Xu, L. Zhang, H. Yang, J. Hou, Y. Wang, Y. Shi, Fate decision of me-
et al. [54] findings that miR-99b, a miR-99a-5p close related family senchymal stem cells: adipocytes or osteoblasts? Cell Death Differ. 23 (7) (2016)
member, has a pro-osteoclastogenesis activity. However, Zhou et al. 1128–1139, https://doi.org/10.1038/cdd.2015.168.
[55] demonstrated that miR-100, other family member, inhibits os- [8] A.M. Silva, S.R. Moura, J.H. Teixeira, M.A. Barbosa, S.G. Santos, M.I. Almeida, Long
noncoding RNAs: a missing link in osteoporosis, Bone Res 7 (2019) 10, https://doi.
teoclast differentiation. Importantly, the nucleotide differences in the org/10.1038/s41413-019-0048-9.
mature sequences from miRNA family members are sufficient to cause [9] M.I. Almeida, R.M. Reis, G.A. Calin, MicroRNA history: discovery, recent applica-
alterations in the targets binding capacity, through modulation of target tions, and next frontiers, Mutat. Res. 717 (1–2) (2011) 1–8, https://doi.org/10.
1016/j.mrfmmm.2011.03.009.
specificity or the strength of repression [56]. [10] G. Papaioannou, F. Mirzamohammadi, T. Kobayashi, MicroRNAs involved in bone
Globally, our results show that miR-99a-5p is a bidirectional reg- formation, Cell. Mol. Life Sci. 71 (24) (2014) 4747–4761, https://doi.org/10.1007/
ulator of bone cellular processes, by reciprocally acting as an inhibitor s00018-014-1700-6.
[11] M.I. Almeida, A.M. Silva, D.M. Vasconcelos, C.R. Almeida, H. Caires, M.T. Pinto,
of osteogenic differentiation and as a promoter of osteoclastogenesis. G.A. Calin, S.G. Santos, M.A. Barbosa, miR-195 in human primary mesenchymal
Furthermore, it also interferes with the intercellular communication stromal/stem cells regulates proliferation, osteogenesis and paracrine effect on
from the bone forming to bone resorbing cells. Taken together, our angiogenesis, Oncotarget 7 (1) (2016) 7–22, https://doi.org/10.18632/oncotarget.
6589.
results support miR-99a-5p as a potential target for the regeneration of [12] D. Li, J. Liu, B. Guo, C. Liang, L. Dang, C. Lu, X. He, H.Y. Cheung, L. Xu, C. Lu, B. He,
bone defects and/or treatment of bone disorders, such as osteoporosis. B. Liu, A.B. Shaikh, F. Li, L. Wang, Z. Yang, D.W. Au, S. Peng, Z. Zhang, B.T. Zhang,
X. Pan, A. Qian, P. Shang, L. Xiao, B. Jiang, C.K. Wong, J. Xu, Z. Bian, Z. Liang,
D.A. Guo, H. Zhu, W. Tan, A. Lu, G. Zhang, Osteoclast-derived exosomal miR-214-
Acknowledgments
3p inhibits osteoblastic bone formation, Nat. Commun. 7 (2016) 10872, https://doi.
org/10.1038/ncomms10872.
This project is supported by Fundação para a Ciência e a Tecnologia [13] J.Y. Krzeszinski, W. Wei, H. Huynh, Z. Jin, X. Wang, T.C. Chang, X.J. Xie, L. He,
(FCT) - in the framework of the project POCI-01-0145-FEDER-031402- L.S. Mangala, G. Lopez-Berestein, A.K. Sood, J.T. Mendell, Y. Wan, miR-34a blocks
osteoporosis and bone metastasis by inhibiting osteoclastogenesis and Tgif2, Nature
R2Bone, under the PORTUGAL 2020 Partnership Agreement, through 512 (7515) (2014) 431–435, https://doi.org/10.1038/nature13375.
ERDF, co-funded by FEDER/FNR, and national funding (through FCT – [14] M. Hadjiargyrou, J. Zhi, D.E. Komatsu, Identification of the microRNA tran-
Fundação para a Ciência e a Tecnologia, I.P., provided by the contract- scriptome during the early phases of mammalian fracture repair, Bone 87 (2016)
78–88, https://doi.org/10.1016/j.bone.2016.03.011.
program and according to numbers 4, 5 and 6 of art. 23 of Law No. 57/ [15] T. Waki, S.Y. Lee, T. Niikura, T. Iwakura, Y. Dogaki, E. Okumachi, K. Oe, R. Kuroda,
2016 of 29 August 2016, as amended by Law No. 57/2017 of 19 July M. Kurosaka, Profiling microRNA expression during fracture healing, BMC
2017). SRM and JB are supported by FCT (SFRH/BD/147229/2019, Musculoskelet. Disord. 17 (2016) 83, https://doi.org/10.1186/s12891-016-0931-0.
[16] L. De-Ugarte, G. Yoskovitz, S. Balcells, R. Guerri-Fernandez, S. Martinez-Diaz,
PD/BD/135490/2018, respectively; BiotechHealth PhD program). The L. Mellibovsky, R. Urreizti, X. Nogues, D. Grinberg, N. Garcia-Giralt, A. Diez-Perez,
mass spectrometry technique was performed at the i3S Proteomics MiRNA profiling of whole trabecular bone: identification of osteoporosis-related
Scientific Platform. This work had support from the Portuguese Mass changes in MiRNAs in human hip bones, BMC Med. Genet. 8 (2015) 75, https://doi.
org/10.1186/s12920-015-0149-2.
Spectrometry Network, integrated in the National Roadmap of Research
[17] C. Seeliger, K. Karpinski, A.T. Haug, H. Vester, A. Schmitt, J.S. Bauer, M. van
Infrastructures of Strategic Relevance (ROTEIRO/0028/2013; LISBOA- Griensven, Five freely circulating miRNAs and bone tissue miRNAs are associated
01-0145-FEDER-022125). with osteoporotic fractures, J. Bone Miner. Res. 29 (8) (2014) 1718–1728, https://
doi.org/10.1002/jbmr.2175.
[18] P. Garmilla-Ezquerra, C. Sanudo, J. Delgado-Calle, M.I. Perez-Nunez, M. Sumillera,
Author contribution J.A. Riancho, Analysis of the bone microRNome in osteoporotic fractures, Calcif.
Tissue Int. 96 (1) (2015) 30–37, https://doi.org/10.1007/s00223-014-9935-7.
SM and MIA: planned all the experiments. SM: performed all the [19] A.M. Silva, M.I. Almeida, J.H. Teixeira, C. Ivan, J. Oliveira, D. Vasconcelos,
N. Neves, C. Ribeiro-Machado, C. Cunha, M.A. Barbosa, G.A. Calin, S.G. Santos,
experiments, data acquisition and wrote the main manuscript text. SM, Profiling the circulating miRnome reveals a temporal regulation of the bone injury
JB, HO: proteomics analysis; SM, JB, JF, MAB, SGS and MIA: discussed response, Theranostics 8 (14) (2018) 3902–3917, https://doi.org/10.7150/thno.
the experiments, reviewed and approved the manuscript. 24444.
[20] J.P. Bras, A.M. Silva, G.A. Calin, M.A. Barbosa, S.G. Santos, M.I. Almeida, miR-195
inhibits macrophages pro-inflammatory profile and impacts the crosstalk with
Declaration of competing interest smooth muscle cells, PLoS One 12 (11) (2017) e0188530, , https://doi.org/10.
1371/journal.pone.0188530.
[21] X. Liang, L. Zhang, S. Wang, Q. Han, R.C. Zhao, Exosomes secreted by mesenchymal
The authors declare no conflict of interest. stem cells promote endothelial cell angiogenesis by transferring miR-125a, J. Cell
Sci. 129 (11) (2016) 2182–2189, https://doi.org/10.1242/jcs.170373.
Appendix A. Supplementary data [22] Y. Tang, L. Zhang, T. Tu, Y. Li, D. Murray, Q. Tu, J.J. Chen, MicroRNA-99a is a
novel regulator of KDM6B-mediated osteogenic differentiation of BMSCs, J. Cell.
Mol. Med. 22 (4) (2018) 2162–2176, https://doi.org/10.1111/jcmm.13490.
Supplementary data to this article can be found online at https:// [23] W. Ding, S. Ding, J. Li, Z. Peng, P. Hu, T. Zhang, L. Pan, Aberrant expression of miR-
doi.org/10.1016/j.bone.2020.115303. 100 in plasma of patients with osteoporosis and its potential diagnostic value, Clin.
Lab. 65 (9) (2019), https://doi.org/10.7754/Clin.Lab.2019.190327.
[24] S. Kelch, E.R. Balmayor, C. Seeliger, H. Vester, J.S. Kirschke, M. van Griensven,
References miRNAs in bone tissue correlate to bone mineral density and circulating miRNAs
are gender independent in osteoporotic patients, Sci. Rep. 7 (1) (2017) 15861, ,
[1] X. Chen, Z. Wang, N. Duan, G. Zhu, E.M. Schwarz, C. Xie, Osteoblast-osteoclast https://doi.org/10.1038/s41598-017-16113-x.
interactions, Connect. Tissue Res. 59 (2) (2018) 99–107, https://doi.org/10.1080/ [25] M. Dominici, K. Le Blanc, I. Mueller, I. Slaper-Cortenbach, F. Marini, D. Krause,
03008207.2017.1290085. R. Deans, A. Keating, D. Prockop, E. Horwitz, Minimal criteria for defining multi-
[2] B.F. Boyce, L. Xing, Functions of RANKL/RANK/OPG in bone modeling and potent mesenchymal stromal cells. The International Society for Cellular Therapy
position statement, Cytotherapy 8 (4) (2006) 315–317, https://doi.org/10.1080/

14
S.R. Moura, et al. Bone 134 (2020) 115303

14653240600855905. Biochem Anal Stud 2 (1) (2017), https://doi.org/10.16966/2576-5833.109.


[26] M.I. Oliveira, S.G. Santos, M.J. Oliveira, A.L. Torres, M.A. Barbosa, Chitosan drives [43] M. Goldberg, D. Septier, A. Oldberg, M.F. Young, L.G. Ameye, Fibromodulin-defi-
anti-inflammatory macrophage polarisation and pro-inflammatory dendritic cell cient mice display impaired collagen fibrillogenesis in predentin as well as altered
stimulation, Eur Cell Mater 24 (2012) 136–152 discussion 152-3 10.22203/ecm. dentin mineralization and enamel formation, J. Histochem. Cytochem. 54 (5)
v024a10. (2006) 525–537, https://doi.org/10.1369/jhc.5A6650.2005.
[27] H. Sudo, H.A. Kodama, Y. Amagai, S. Yamamoto, S. Kasai, In vitro differentiation [44] C.J. Li, Y. Xiao, M. Yang, T. Su, X. Sun, Q. Guo, Y. Huang, X.H. Luo, Long noncoding
and calcification in a new clonal osteogenic cell line derived from newborn mouse RNA Bmncr regulates mesenchymal stem cell fate during skeletal aging, J. Clin.
calvaria, J. Cell Biol. 96 (1) (1983) 191–198, https://doi.org/10.1083/jcb.96.1. Invest. 128 (12) (2018) 5251–5266, https://doi.org/10.1172/JCI99044.
191. [45] X. Ren, Q. Zhou, D. Foulad, A.S. Tiffany, M.J. Dewey, D. Bischoff, T.A. Miller,
[28] P. Collin-Osdoby, P. Osdoby, RANKL-mediated osteoclast formation from murine R.R. Reid, T.C. He, D.T. Yamaguchi, B.A.C. Harley, J.C. Lee, Osteoprotegerin re-
RAW 264.7 cells, Methods Mol. Biol. 816 (2012) 187–202, https://doi.org/10. duces osteoclast resorption activity without affecting osteogenesis on nanoparti-
1007/978-1-61779-415-5_13. culate mineralized collagen scaffolds, Sci. Adv. 5 (6) (2019) eaaw4991, , https://
[29] A.H. Lourenco, A.L. Torres, D.P. Vasconcelos, C. Ribeiro-Machado, J.N. Barbosa, doi.org/10.1126/sciadv.aaw4991.
M.A. Barbosa, C.C. Barrias, C.C. Ribeiro, Osteogenic, anti-osteoclastogenic and [46] W.S. Simonet, D.L. Lacey, C.R. Dunstan, M. Kelley, M.S. Chang, R. Luthy,
immunomodulatory properties of a strontium-releasing hybrid scaffold for bone H.Q. Nguyen, S. Wooden, L. Bennett, T. Boone, G. Shimamoto, M. DeRose, R. Elliott,
repair, Mater. Sci. Eng. C Mater. Biol. Appl. 99 (2019) 1289–1303, https://doi.org/ A. Colombero, H.L. Tan, G. Trail, J. Sullivan, E. Davy, N. Bucay, L. Renshaw-Gegg,
10.1016/j.msec.2019.02.053. T.M. Hughes, D. Hill, W. Pattison, P. Campbell, S. Sander, G. Van, J. Tarpley,
[30] S. Yan, L. Miao, Y. Lu, L. Wang, Sirtuin 1 inhibits TNF-alpha-mediated osteoclas- P. Derby, R. Lee, W.J. Boyle, Osteoprotegerin: a novel secreted protein involved in
togenesis of bone marrow-derived macrophages through both ROS generation and the regulation of bone density, Cell 89 (2) (1997) 309–319, https://doi.org/10.
TRPV1 activation, Mol. Cell. Biochem. 455 (1–2) (2019) 135–145, https://doi.org/ 1016/s0092-8674(00)80209-3.
10.1007/s11010-018-3477-7. [47] H. Brandstrom, T. Bjorkman, O. Ljunggren, Regulation of osteoprotegerin secretion
[31] A.L. Torres, S.G. Santos, M.I. Oliveira, M.A. Barbosa, Fibrinogen promotes resorp- from primary cultures of human bone marrow stromal cells, Biochem. Biophys. Res.
tion of chitosan by human osteoclasts, Acta Biomater. 9 (5) (2013) 6553–6562, Commun. 280 (3) (2001) 831–835, https://doi.org/10.1006/bbrc.2000.4223.
https://doi.org/10.1016/j.actbio.2013.01.015. [48] H. Min, S. Morony, I. Sarosi, C.R. Dunstan, C. Capparelli, S. Scully, G. Van,
[32] C.R. Gardner, Morphological analysis of osteoclastogenesis induced by RANKL in S. Kaufman, P.J. Kostenuik, D.L. Lacey, W.J. Boyle, W.S. Simonet, Osteoprotegerin
mouse bone marrow cell cultures, Cell Biol. Int. 31 (7) (2007) 672–682, https://doi. reverses osteoporosis by inhibiting endosteal osteoclasts and prevents vascular
org/10.1016/j.cellbi.2006.12.008. calcification by blocking a process resembling osteoclastogenesis, J. Exp. Med. 192
[33] S.A. Bustin, V. Benes, J.A. Garson, J. Hellemans, J. Huggett, M. Kubista, R. Mueller, (4) (2000) 463–474, https://doi.org/10.1084/jem.192.4.463.
T. Nolan, M.W. Pfaffl, G.L. Shipley, J. Vandesompele, C.T. Wittwer, The MIQE [49] X. Liu, C. Bao, H.H.K. Xu, J. Pan, J. Hu, P. Wang, E. Luo, Osteoprotegerin gene-
guidelines: minimum information for publication of quantitative real-time PCR modified BMSCs with hydroxyapatite scaffold for treating critical-sized mandibular
experiments, Clin. Chem. 55 (4) (2009) 611–622, https://doi.org/10.1373/ defects in ovariectomized osteoporotic rats, Acta Biomater. 42 (2016) 378–388,
clinchem.2008.112797. https://doi.org/10.1016/j.actbio.2016.06.019.
[34] H. Dweep, C. Sticht, P. Pandey, N. Gretz, miRWalk–database: prediction of possible [50] A. Pinheiro, A.M. Silva, J.H. Teixeira, R.M. Goncalves, M.I. Almeida, M.A. Barbosa,
miRNA binding sites by “walking” the genes of three genomes, J. Biomed. Inform. S.G. Santos, Extracellular vesicles: intelligent delivery strategies for therapeutic
44 (5) (2011) 839–847, https://doi.org/10.1016/j.jbi.2011.05.002. applications, J. Control. Release 289 (2018) 56–69, https://doi.org/10.1016/j.
[35] T.F. Tsai, J.F. Lin, K.Y. Chou, Y.C. Lin, H.E. Chen, T.I. Hwang, miR-99a-5p acts as jconrel.2018.09.019.
tumor suppressor via targeting to mTOR and enhances RAD001-induced apoptosis [51] A.M. Silva, J.H. Teixeira, M.I. Almeida, R.M. Goncalves, M.A. Barbosa, S.G. Santos,
in human urinary bladder urothelial carcinoma cells, Onco Targets Ther 11 (2018) Extracellular vesicles: Immunomodulatory messengers in the context of tissue re-
239–252, https://doi.org/10.2147/OTT.S114276. pair/regeneration, Eur. J. Pharm. Sci. 98 (2017) 86–95, https://doi.org/10.1016/j.
[36] L.L. Mei, Y.T. Qiu, M.B. Huang, W.J. Wang, J. Bai, Z.Z. Shi, MiR-99a suppresses ejps.2016.09.017.
proliferation, migration and invasion of esophageal squamous cell carcinoma cells [52] A. Turchinovich, L. Weiz, A. Langheinz, B. Burwinkel, Characterization of extra-
through inhibiting the IGF1R signaling pathway, Cancer Biomark 20 (4) (2017) cellular circulating microRNA, Nucleic Acids Res. 39 (16) (2011) 7223–7233,
527–537, https://doi.org/10.3233/CBM-170345. https://doi.org/10.1093/nar/gkr254.
[37] L. Cui, H. Zhou, H. Zhao, Y. Zhou, R. Xu, X. Xu, L. Zheng, Z. Xue, W. Xia, B. Zhang, [53] L. de la Rica, A. Garcia-Gomez, N.R. Comet, J. Rodriguez-Ubreva, L. Ciudad,
T. Ding, Y. Cao, Z. Tian, Q. Shi, X. He, MicroRNA-99a induces G1-phase cell cycle R. Vento-Tormo, C. Company, D. Alvarez-Errico, M. Garcia, C. Gomez-Vaquero,
arrest and suppresses tumorigenicity in renal cell carcinoma, BMC Cancer 12 (2012) E. Ballestar, NF-kappaB-direct activation of microRNAs with repressive effects on
546, https://doi.org/10.1186/1471-2407-12-546. monocyte-specific genes is critical for osteoclast differentiation, Genome Biol. 16
[38] H. Qin, W. Liu, MicroRNA-99a-5p suppresses breast cancer progression and cell- (2015) 2, https://doi.org/10.1186/s13059-014-0561-5.
cycle pathway through downregulating CDC25A, J. Cell. Physiol. 234 (4) (2019) [54] T. Franceschetti, N.S. Dole, C.B. Kessler, S.K. Lee, A.M. Delany, Pathway analysis of
3526–3537, https://doi.org/10.1002/jcp.26906. microRNA expression profile during murine osteoclastogenesis, PLoS One 9 (9)
[39] L.F. Gulyaeva, N.E. Kushlinskiy, Regulatory mechanisms of microRNA expression, (2014) e107262, , https://doi.org/10.1371/journal.pone.0107262.
J. Transl. Med. 14 (1) (2016) 143, https://doi.org/10.1186/s12967-016-0893-x. [55] L. Zhou, H.Y. Song, L.L. Gao, L.Y. Yang, S. Mu, Q. Fu, MicroRNA1005p inhibits
[40] A. Hamann, A. Nguyen, A.K. Pannier, Nucleic acid delivery to mesenchymal stem osteoclastogenesis and bone resorption by regulating fibroblast growth factor 21,
cells: a review of nonviral methods and applications, J. Biol. Eng. 13 (2019) 7, Int. J. Mol. Med. 43 (2) (2019) 727–738, https://doi.org/10.3892/ijmm.2018.
https://doi.org/10.1186/s13036-019-0140-0. 4017.
[41] A.Y. Ng, C. Tu, S. Shen, D. Xu, M.J. Oursler, J. Qu, S. Yang, Comparative char- [56] J.M. Wolter, H.H. Le, A. Linse, V.A. Godlove, T.D. Nguyen, K. Kotagama, A. Lynch,
acterization of osteoclasts derived from murine bone marrow macrophages and A. Rawls, M. Mangone, Evolutionary patterns of metazoan microRNAs reveal tar-
RAW 264.7 cells using quantitative proteomics, JBMR Plus 2 (6) (2018) 328–340, geting principles in the let-7 and miR-10 families, Genome Res. 27 (1) (2017)
https://doi.org/10.1002/jbm4.10058. 53–63, https://doi.org/10.1101/gr.209361.116.
[42] J. Nguyen, A. Nohe, Factors that affect the osteoclastogenesis of RAW264.7 cells, J

15

You might also like