Legumain Regulates Differentiation Fate of Human Bone Marrow Stromal Cells and Is Altered in Postmenopausal Osteoporosis

Abbas Jafari, Diyako Qanie, Thomas L Andersen, Yuxi Zhang, Li Chen, Benno Postert, Stuart Parsons, Nicholas Ditzel, Sundeep Khosla, Harald Thidemann Johansen, Per Kjærsgaard-Andersen, Jean-Marie Delaisse, Basem M Abdallah, Daniel Hesselson, Rigmor Solberg, Moustapha Kassem, Abbas Jafari, Diyako Qanie, Thomas L Andersen, Yuxi Zhang, Li Chen, Benno Postert, Stuart Parsons, Nicholas Ditzel, Sundeep Khosla, Harald Thidemann Johansen, Per Kjærsgaard-Andersen, Jean-Marie Delaisse, Basem M Abdallah, Daniel Hesselson, Rigmor Solberg, Moustapha Kassem

Abstract

Secreted factors are a key component of stem cell niche and their dysregulation compromises stem cell function. Legumain is a secreted cysteine protease involved in diverse biological processes. Here, we demonstrate that legumain regulates lineage commitment of human bone marrow stromal cells and that its expression level and cellular localization are altered in postmenopausal osteoporotic patients. As shown by genetic and pharmacological manipulation, legumain inhibited osteoblast (OB) differentiation and in vivo bone formation through degradation of the bone matrix protein fibronectin. In addition, genetic ablation or pharmacological inhibition of legumain activity led to precocious OB differentiation and increased vertebral mineralization in zebrafish. Finally, we show that localized increased expression of legumain in bone marrow adipocytes was inversely correlated with adjacent trabecular bone mass in a cohort of patients with postmenopausal osteoporosis. Our data suggest that altered proteolytic activity of legumain in the bone microenvironment contributes to decreased bone mass in postmenopausal osteoporosis.

Keywords: adipocyte; bone marrow stromal cells; differentiation; extracellular matrix; fibronectin; legumain; mesenchymal stem cell; osteoblast; osteoporosis; proliferation.

Copyright © 2017 The Authors. Published by Elsevier Inc. All rights reserved.

Figures

Figure 1
Figure 1
Regulation of Legumain Expression during In Vitro and In Vivo Differentiation of Human Bone Marrow Stromal Cells (A and B) Immunohistochemical (A) and RNA in situ hybridization (B) analysis of legumain expression and localization in normal human iliac crest bone biopsies. n = 11 donors. Scale bar, 50 μm. Red arrows, canopy cells; black arrows, reversal cells; arrow heads, osteoclasts; v, vessel. (C) qRT-PCR analysis of LGMN expression during osteoblast (OB) differentiation of hBMSCs at 6, 12, and 18 days (D6–D18) after start of differentiation (day 0, D0). Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (D) Western blot analysis of legumain expression in cell lysates from hBMSC cultures during OB differentiation. (E) Quantification of the mature legumain (36 kDa) band intensity. Arbitrary units (ARBU). Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (F) Quantification of legumain activity in cell lysates from hBMSCs during OB differentiation. Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, two-tailed unpaired Student’s t test. (G) qRT-PCR analysis of LGMN expression during adipocyte (AD) differentiation of hBMSCs. Data represent mean ± SD from three independent experiments. ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (H) Western blot analysis of legumain expression in cell lysates from hBMSC cultures during AD differentiation. (I) Quantification of the mature legumain (36 kDa) band intensity. Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, two-tailed unpaired Student’s t test.
Figure 2
Figure 2
Legumain Knockdown Enhanced Osteoblast Differentiation and In Vivo Bone Formation and Inhibited Adipocyte Differentiation of Human Bone Marrow Stromal Cells hBMSCs were stably transfected with control (shCtrl) or LGMN shRNA (shLGMN). (A) qRT-PCR analysis of LGMN expression. Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, two-tailed unpaired Student’s t test. (B) Western blot analysis of legumain and GAPDH control. Data represent three independent experiments. (C) Quantification of legumain activity. Data represent mean ± SD from three independent experiments. ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (D) Quantification of alkaline phosphatase (ALP) activity in the presence of standard culture medium (SCM) or osteoblast induction medium (OIM) (day 6). Data represent mean ± SD from three independent experiments. p > 0.05, two-tailed unpaired Student’s t test. (E) qRT-PCR gene expression analysis of the early (ALP, Col1a1) and late (BGLAP, IBSP) OB marker genes. Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (F) Quantification of alizarin red staining at day 12. Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, two-tailed unpaired Student’s t test. (G and H) Quantification of accumulated lipid droplets in the presence of AD induction medium using oil red O staining (day 12). Scale bar, 150 μm, Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, two-tailed unpaired Student’s t test. (I) qRT-PCR gene expression analysis of the AD marker genes PPARG2, FABP4, LPL, and ADIPOQ. Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, two-tailed unpaired Student’s t test. (J) Histological analysis of in vivo bone formation by hBMSCs stably transfected with non-targeting control shRNA (shCtrl) or LGMN shRNA (shLGMN), 8 weeks after implantation in immune-deficient mice. Arrows, hydroxyapatite; arrow heads, bone. Scale bars: top panels, 500 μm; bottom panels, 250 μm. (K) Quantification of the heterotopic bone formation, n = 4 implants for each cell type, Data represent mean ± SEM. ∗p ≤ 0.05, Mann-Whitney test. (L) Human-specific vimentin staining of shLGMN implants. Arrows, hydroxyapatite; arrow heads, bone. Scale bars: top panels, 500 μm; bottom panels, 250 μm. ab, antibody. See also Figure S1.
Figure 3
Figure 3
Legumain Overexpression Inhibited Osteoblast and Enhanced Adipocyte Differentiation of Human Bone Marrow Stromal Cells See also Figure S1. (A–C) Legumain (LGMN)-transduced hBMSCs was established using a retroviral transduction system and the successful overexpression of legumain was confirmed using (A) qRT-PCR analysis of LGMN mRNA expression, (B) western blot analysis of legumain in cell lysate, and (C) quantification of legumain activity. Data represent mean ± SD from three independent experiments. ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (D–G) Secretion of legumain in the conditioned medium (CM) was evaluated using (D) ELISA measurement (data represent mean from three technical replicates) and (E) western blot analysis of legumain in the CM from hBMSC-LGMN-overexpressing cell line (LGMN-CM) (data represent three independent experiments). To assess the effects of legumain on OB differentiation, control hBMSCs containing empty vector (E.V.) and hBMSC-LGMN cell lines were cultured in OB induction medium, and expressions of OB marker genes were analyzed using qRT-PCR (F) and mineralized matrix formation was determined by quantification of eluted alizarin red staining (G). Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (H and I) To assess the effect of legumain on AD differentiation, control hBMSCs and hBMSC-LGMN cell lines were cultured in AD induction medium, and (H), (I) accumulation of lipid droplets was measured by quantification of the eluted oil red O staining, ∗p ≤ 0.05, two-tailed unpaired Student’s t test. Scale bar, 150 μm. (J) Expressions of AD marker genes were measured by qRT-PCR (day 7). Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, two-tailed unpaired Student’s t test. E.V., empty vector; LGMN, legumain-overexpressing hBMSCs.
Figure 4
Figure 4
Legumain Degrades Fibronectin in Human Bone Marrow Stromal Cell cultures (A) Western blot analysis of legumain and fibronectin in cell lysates of hBMSC lines with stable knockdown of legumain (shLGMN) during ex vivo OB differentiation (day 0–7). (B and C) Quantification of protein band intensities in (A). ∗p ≤ 0.05, ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (D) Western blot analysis of legumain and fibronectin in cell lysates of hBMSC lines with stable overexpression of legumain (LGMN) during ex vivo OB differentiation (day 0–7). (E and F) Quantification of protein band intensities in (D). Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (G) Western blot analysis of human fibronectin degradation by purified legumain from bovine kidneys (bLeg; 10:1 w/w; control) and the cell lysates from hBMSCs stably transfected with E.V. or legumain (LGMN; legumain overexpression). (H) Quantitation of mineralized matrix formation on day 15 of OB differentiation, in the presence of siRNA against fibronectin (siFN) and non-targeting control siRNA (siCtrl). Data represent mean ± SD from three independent experiments. ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. (I and J) Effect of different ECM proteins (gelatin, collagen 1, or fibronectin) on mineralized matrix formation by hBMSCs visualized by alizarin red staining and quantification. Data represent mean ± SD from three independent experiments. ∗p ≤ 0.05, ∗∗p ≤ 0.01, two-tailed unpaired Student’s t test. See also Figure S2.
Figure 5
Figure 5
Legumain Inhibits OB Differentiation and Bone Mineralization In Vivo (A) Conservation of zebrafish lgmn. I, amino acid identity; S, amino acid similarity. (B) High-resolution melting analysis of pooled lgmn-TALEN and control-injected animals at 3 days post-fertilization (dpf). (C) qRT-PCR analysis of OB and AD marker genes at 5 dpf. Data represent mean ± SEM of three pools of ten animals. ∗p ≤ 0.05, two-tailed unpaired Student’s t test. (D and E) Control and lgmn mutant animals at 7 dpf. Scale bars, 500 μM. (D) Bright field and (E) fluorescent alizarin red staining. (E′ and E″) Higher-magnification images of boxed regions. Arrowheads, whole vertebrae stained; arrows, partial vertebrae stained. Scale bars, 100 μM. (F) Number of alizarin-red-stained vertebrae in lgmn and control animals at 7 dpf. Data represent mean ± SEM, n > 30 for each group. ∗p ≤ 0.05, ∗∗∗p ≤ 0.005 two-tailed unpaired Student’s t test. (G) Animals were treated with SD-134 (500 μM) or 1% DMSO from 3 to 7 dpf. Number of alizarin-red-stained vertebrae at 7 dpf. Data represent mean ± SEM, n = 12 for each group. ∗∗∗p ≤ 0.005 two-tailed unpaired Student’s t test.
Figure 6
Figure 6
Effect of Aging and Osteoporosis on Legumain Protein Levels in Human Serum and Bone Microenvironment (A) Serum legumain levels in 89 women, aged 48–87 years, p GAPDH in cell lysates from primary hBMSC cultures established from bone marrow aspirates of osteoporotic patients (n = 5) and age-matched controls (n = 3). (C) Quantification of western blot band intensities normalized to GAPDH. Data represent mean ± SD from three independent experiments. ∗∗∗p ≤ 0.005, two-tailed unpaired Student’s t test. (D) Upper panel: legumain staining of bone biopsies from postmenopausal osteoporotic patients (n = 11) and age-matched control individuals (n = 13). Scale bars, 1 mm. Lower panel: higher-magnification images of the boxed regions. Scale bars, 50 μm. (E) Correlation of trabecular bone area with legumain expression by adipocytes in biopsies from postmenopausal osteoporotic patients (n = 80 random regions of interest [ROI] = 1 mm2), p = 0.004 using Spearman correlation analysis. (F) Proposed mode-of-action of legumain for regulation of hBMSC lineage commitment.

References

    1. Abdallah B.M., Kassem M. Human mesenchymal stem cells: from basic biology to clinical applications. Gene Ther. 2008;15:109–116.
    1. Abdallah B.M., Haack-Sorensen M., Burns J.S., Elsnab B., Jakob F., Hokland P., Kassem M. Maintenance of differentiation potential of human bone marrow mesenchymal stem cells immortalized by human telomerase reverse transcriptase gene despite [corrected] extensive proliferation. Biochem. Biophys. Res. Commun. 2005;326:527–538.
    1. Abdallah B.M., Ditzel N., Kassem M. Assessment of bone formation capacity using in vivo transplantation assays: procedure and tissue analysis. Methods Mol. Biol. 2008;455:89–100.
    1. Andersen T.L., Abdelgawad M.E., Kristensen H.B., Hauge E.M., Rolighed L., Bollerslev J., Kjaersgaard-Andersen P., Delaisse J.M. Understanding coupling between bone resorption and formation: are reversal cells the missing link? Am. J. Pathol. 2013;183:235–246.
    1. Andrade V., Guerra M., Jardim C., Melo F., Silva W., Ortega J.M., Robert M., Nathanson M.H., Leite F. Nucleoplasmic calcium regulates cell proliferation through legumain. J. Hepatol. 2011;55:626–635.
    1. Antras J., Hilliou F., Redziniak G., Pairault J. Decreased biosynthesis of actin and cellular fibronectin during adipose conversion of 3T3-F442A cells. Reorganization of the cytoarchitecture and extracellular matrix fibronectin. Biol. Cell. 1989;66:247–254.
    1. Ashley S.L., Xia M., Murray S., O'Dwyer D.N., Grant E., White E.S., Flaherty K.R., Martinez F.J., Moore B.B. Six-SOMAmer index relating to immune, protease and angiogenic functions predicts progression in IPF. PLoS One. 2016;11:e0159878.
    1. Bey E., Prat M., Duhamel P., Benderitter M., Brachet M., Trompier F., Battaglini P., Ernou I., Boutin L., Gourven M. Emerging therapy for improving wound repair of severe radiation burns using local bone marrow-derived stem cell administrations. Wound Repair Regen. 2010;18:50–58.
    1. Brunner M., Millon-Fremillon A., Chevalier G., Nakchbandi I.A., Mosher D., Block M.R., Albiges-Rizo C., Bouvard D. Osteoblast mineralization requires beta1 integrin/ICAP-1-dependent fibronectin deposition. J. Cell Biol. 2011;194:307–322.
    1. Charatcharoenwitthaya N., Khosla S., Atkinson E.J., McCready L.K., Riggs B.L. Effect of blockade of TNF-alpha and interleukin-1 action on bone resorption in early postmenopausal women. J. Bone Miner Res. 2007;22:724–729.
    1. Chen J.M., Fortunato M., Stevens R.A., Barrett A.J. Activation of progelatinase A by mammalian legumain, a recently discovered cysteine proteinase. Biol. Chem. 2001;382:777–783.
    1. Choi S.J., Reddy S.V., Devlin R.D., Menaa C., Chung H., Boyce B.F., Roodman G.D. Identification of human asparaginyl endopeptidase (legumain) as an inhibitor of osteoclast formation and bone resorption. J. Biol. Chem. 1999;274:27747–27753.
    1. Choi S.J., Kurihara N., Oba Y., Roodman G.D. Osteoclast inhibitory peptide 2 inhibits osteoclast formation via its C-terminal fragment. J. Bone Miner Res. 2001;16:1804–1811.
    1. Clerin V., Shih H.H., Deng N., Hebert G., Resmini C., Shields K.M., Feldman J.L., Winkler A., Albert L., Maganti V. Expression of the cysteine protease legumain in vascular lesions and functional implications in atherogenesis. Atherosclerosis. 2008;201:53–66.
    1. Compston J. Osteoporosis: social and economic impact. Radiol. Clin. North Am. 2010;48:477–482.
    1. Dall E., Brandstetter H. Mechanistic and structural studies on legumain explain its zymogenicity, distinct activation pathways, and regulation. Proc. Natl. Acad. Sci. USA. 2013;110:10940–10945.
    1. Dall E., Fegg J.C., Briza P., Brandstetter H. Structure and mechanism of an aspartimide-dependent peptide ligase in human legumain. Angew. Chem. Int. Ed. Engl. 2015;54:2917–2921.
    1. Delaisse J.M. The reversal phase of the bone-remodeling cycle: cellular prerequisites for coupling resorption and formation. Bonekey Rep. 2014;3:561.
    1. Deryugina E.I., Quigley J.P. Matrix metalloproteinases and tumor metastasis. Cancer Metastasis Rev. 2006;25:9–34.
    1. Detre S., Saclani Jotti G., Dowsett M. A “quickscore” method for immunohistochemical semiquantitation: validation for oestrogen receptor in breast carcinomas. J. Clin. Pathol. 1995;48:876–878.
    1. Ewald S.E., Lee B.L., Lau L., Wickliffe K.E., Shi G.P., Chapman H.A., Barton G.M. The ectodomain of Toll-like receptor 9 is cleaved to generate a functional receptor. Nature. 2008;456:658–662.
    1. Ewald S.E., Engel A., Lee J., Wang M., Bogyo M., Barton G.M. Nucleic acid recognition by Toll-like receptors is coupled to stepwise processing by cathepsins and asparagine endopeptidase. J. Exp. Med. 2011;208:643–651.
    1. Fisher S., Jagadeeswaran P., Halpern M.E. Radiographic analysis of zebrafish skeletal defects. Dev. Biol. 2003;264:64–76.
    1. Grimes D.T., Boswell C.W., Morante N.F., Henkelman R.M., Burdine R.D., Ciruna B. Zebrafish models of idiopathic scoliosis link cerebrospinal fluid flow defects to spine curvature. Science. 2016;352:1341–1344.
    1. Guo P., Zhu Z., Sun Z., Wang Z., Zheng X., Xu H. Expression of legumain correlates with prognosis and metastasis in gastric carcinoma. PLoS One. 2013;8:e73090.
    1. Halper J., Kjaer M. Basic components of connective tissues and extracellular matrix: elastin, fibrillin, fibulins, fibrinogen, fibronectin, laminin, tenascins and thrombospondins. Adv. Exp. Med. Biol. 2014;802:31–47.
    1. Hammond J.B., Kruger N.J. The Bradford method for protein quantitation. Methods Mol. Biol. 1988;3:25–32.
    1. Hare J.M., Traverse J.H., Henry T.D., Dib N., Strumpf R.K., Schulman S.P., Gerstenblith G., DeMaria A.N., Denktas A.E., Gammon R.S. A randomized, double-blind, placebo-controlled, dose-escalation study of intravenous adult human mesenchymal stem cells (prochymal) after acute myocardial infarction. J. Am. Coll. Cardiol. 2009;54:2277–2286.
    1. Hayes M., Gao X., Yu L.X., Paria N., Henkelman R.M., Wise C.A., Ciruna B. ptk7 mutant zebrafish models of congenital and idiopathic scoliosis implicate dysregulated Wnt signalling in disease. Nat. Commun. 2014;5:4777.
    1. Hernigou P., Poignard A., Beaujean F., Rouard H. Percutaneous autologous bone-marrow grafting for nonunions. Influence of the number and concentration of progenitor cells. J. Bone Joint Surg. Am. 2005;87:1430–1437.
    1. Hoshiba T., Kawazoe N., Chen G. The balance of osteogenic and adipogenic differentiation in human mesenchymal stem cells by matrices that mimic stepwise tissue development. Biomaterials. 2012;33:2025–2031.
    1. Jafari A., Siersbaek M.S., Chen L., Qanie D., Zaher W., Abdallah B.M., Kassem M. Pharmacological inhibition of protein kinase G1 enhances bone formation by human skeletal stem cells through activation of RhoA-Akt signaling. Stem Cells. 2015;33:2219–2231.
    1. Johansen H.T., Knight C.G., Barrett A.J. Colorimetric and fluorimetric microplate assays for legumain and a staining reaction for detection of the enzyme after electrophoresis. Anal. Biochem. 1999;273:278–283.
    1. Kimmel C.B., Ballard W.W., Kimmel S.R., Ullmann B., Schilling T.F. Stages of embryonic development of the zebrafish. Dev. Dyn. 1995;203:253–310.
    1. Kitamoto S., Sukhova G.K., Sun J., Yang M., Libby P., Love V., Duramad P., Sun C., Zhang Y., Yang X. Cathepsin L deficiency reduces diet-induced atherosclerosis in low-density lipoprotein receptor-knockout mice. Circulation. 2007;115:2065–2075.
    1. Kristensen L.P., Chen L., Nielsen M.O., Qanie D.W., Kratchmarova I., Kassem M., Andersen J.S. Temporal profiling and pulsed SILAC labeling identify novel secreted proteins during ex vivo osteoblast differentiation of human stromal stem cells. Mol. Cell. Proteomics. 2012;11:989–1007.
    1. Kristensen H.B., Andersen T.L., Marcussen N., Rolighed L., Delaisse J.M. Osteoblast recruitment routes in human cancellous bone remodeling. Am. J. Pathol. 2014;184:778–789.
    1. Lee J., Bogyo M. Synthesis and evaluation of aza-peptidyl inhibitors of the lysosomal asparaginyl endopeptidase, legumain. Bioorg. Med. Chem. Lett. 2012;22:1340–1343.
    1. Lee R.H., Kim B., Choi I., Kim H., Choi H.S., Suh K., Bae Y.C., Jung J.S. Characterization and expression analysis of mesenchymal stem cells from human bone marrow and adipose tissue. Cell Physiol. Biochem. 2004;14:311–324.
    1. Li N., Felber K., Elks P., Croucher P., Roehl H.H. Tracking gene expression during zebrafish osteoblast differentiation. Dev. Dyn. 2009;238:459–466.
    1. Lin Y., Qiu Y., Xu C., Liu Q., Peng B., Kaufmann G.F., Chen X., Lan B., Wei C., Lu D. Functional role of asparaginyl endopeptidase ubiquitination by TRAF6 in tumor invasion and metastasis. J. Natl. Cancer Inst. 2014;106:dju012.
    1. Linsley C., Wu B., Tawil B. The effect of fibrinogen, collagen type I, and fibronectin on mesenchymal stem cell growth and differentiation into osteoblasts. Tissue Eng. Part A. 2013;19:1416–1423.
    1. Liu C., Sun C., Huang H., Janda K., Edgington T. Overexpression of legumain in tumors is significant for invasion/metastasis and a candidate enzymatic target for prodrug therapy. Cancer Res. 2003;63:2957–2964.
    1. Lunde N.N., Holm S., Dahl T.B., Elyouncha I., Sporsheim B., Gregersen I., Abbas A., Skjelland M., Espevik T., Solberg R. Increased levels of legumain in plasma and plaques from patients with carotid atherosclerosis. Atherosclerosis. 2016
    1. Manoury B., Hewitt E.W., Morrice N., Dando P.M., Barrett A.J., Watts C. An asparaginyl endopeptidase processes a microbial antigen for class II MHC presentation. Nature. 1998;396:695–699.
    1. Mathews S., Bhonde R., Gupta P.K., Totey S. Extracellular matrix protein mediated regulation of the osteoblast differentiation of bone marrow derived human mesenchymal stem cells. Differentiation. 2012;84:185–192.
    1. Mattock K.L., Gough P.J., Humphries J., Burnand K., Patel L., Suckling K.E., Cuello F., Watts C., Gautel M., Avkiran M. Legumain and cathepsin-L expression in human unstable carotid plaque. Atherosclerosis. 2010;208:83–89.
    1. Miller G., Matthews S.P., Reinheckel T., Fleming S., Watts C. Asparagine endopeptidase is required for normal kidney physiology and homeostasis. FASEB J. 2011;25:1606–1617.
    1. Morita Y., Araki H., Sugimoto T., Takeuchi K., Yamane T., Maeda T., Yamamoto Y., Nishi K., Asano M., Shirahama-Noda K. Legumain/asparaginyl endopeptidase controls extracellular matrix remodeling through the degradation of fibronectin in mouse renal proximal tubular cells. FEBS Lett. 2007;581:1417–1424.
    1. Moursi A.M., Globus R.K., Damsky C.H. Interactions between integrin receptors and fibronectin are required for calvarial osteoblast differentiation in vitro. J. Cell Sci. 1997;110:2187–2196.
    1. Papaspyridonos M., Smith A., Burnand K.G., Taylor P., Padayachee S., Suckling K.E., James C.H., Greaves D.R., Patel L. Novel candidate genes in unstable areas of human atherosclerotic plaques. Arterioscler. Thromb. Vasc. Biol. 2006;26:1837–1844.
    1. Rodriguez Fernandez J.L., Ben-Ze'ev A. Regulation of fibronectin, integrin and cytoskeleton expression in differentiating adipocytes: inhibition by extracellular matrix and polylysine. Differentiation. 1989;42:65–74.
    1. Sakuma T., Ochiai H., Kaneko T., Mashimo T., Tokumasu D., Sakane Y., Suzuki K., Miyamoto T., Sakamoto N., Matsuura S. Repeating pattern of non-RVD variations in DNA-binding modules enhances TALEN activity. Sci. Rep. 2013;3:3379.
    1. Schuck S., Manninen A., Honsho M., Fullekrug J., Simons K. Generation of single and double knockdowns in polarized epithelial cells by retrovirus-mediated RNA interference. Proc. Natl. Acad. Sci. USA. 2004;101:4912–4917.
    1. Sepulveda F.E., Maschalidi S., Colisson R., Heslop L., Ghirelli C., Sakka E., Lennon-Dumenil A.M., Amigorena S., Cabanie L., Manoury B. Critical role for asparagine endopeptidase in endocytic Toll-like receptor signaling in dendritic cells. Immunity. 2009;31:737–748.
    1. Shah A.N., Davey C.F., Whitebirch A.C., Miller A.C., Moens C.B. Rapid reverse genetic screening using CRISPR in zebrafish. Nat. Methods. 2015;12:535–540.
    1. Simonsen J.L., Rosada C., Serakinci N., Justesen J., Stenderup K., Rattan S.I., Jensen T.G., Kassem M. Telomerase expression extends the proliferative life-span and maintains the osteogenic potential of human bone marrow stromal cells. Nat. Biotechnol. 2002;20:592–596.
    1. Smith R., Johansen H.T., Nilsen H., Haugen M.H., Pettersen S.J., Maelandsmo G.M., Abrahamson M., Solberg R. Intra- and extracellular regulation of activity and processing of legumain by cystatin E/M. Biochimie. 2012;94:2590–2599.
    1. Solberg R., Smith R., Almlof M., Tewolde E., Nilsen H., Johansen H.T. Legumain expression, activity and secretion are increased during monocyte-to-macrophage differentiation and inhibited by atorvastatin. Biol. Chem. 2015;396:71–80.
    1. Spiegelman B.M., Ginty C.A. Fibronectin modulation of cell shape and lipogenic gene expression in 3T3-adipocytes. Cell. 1983;35:657–666.
    1. Wang L., Chen S., Zhang M., Li N., Chen Y., Su W., Liu Y., Lu D., Li S., Yang Y. Legumain: a biomarker for diagnosis and prognosis of human ovarian cancer. J. Cell Biochem. 2012;113:2679–2686.
    1. Westerfield M. University of Oregon Press; 2007. The Zebrafish Book: A Guide for the Laboratory Use of Zebrafish (Danio Rerio)
    1. Wu M., Shao G.R., Zhang F.X., Wu W.X., Xu P., Ruan Z.M. Legumain protein as a potential predictive biomarker for Asian patients with breast carcinoma. Asian Pac. J. Cancer Prev. 2014;15:10773–10777.
    1. Yamout B., Hourani R., Salti H., Barada W., El-Hajj T., Al-Kutoubi A., Herlopian A., Baz E.K., Mahfouz R., Khalil-Hamdan R. Bone marrow mesenchymal stem cell transplantation in patients with multiple sclerosis: a pilot study. J. Neuroimmunol. 2010;227:185–189.

Source: PubMed

3
Prenumerera