Membrane-stabilizing copolymers confer marked protection to dystrophic skeletal muscle in vivo

Evelyne M Houang, Karen J Haman, Antonio Filareto, Rita C Perlingeiro, Frank S Bates, Dawn A Lowe, Joseph M Metzger, Evelyne M Houang, Karen J Haman, Antonio Filareto, Rita C Perlingeiro, Frank S Bates, Dawn A Lowe, Joseph M Metzger

Abstract

Duchenne muscular dystrophy (DMD) is a fatal disease of striated muscle deterioration. A unique therapeutic approach for DMD is the use of synthetic membrane stabilizers to protect the fragile dystrophic sarcolemma against contraction-induced mechanical stress. Block copolymer-based membrane stabilizer poloxamer 188 (P188) has been shown to protect the dystrophic myocardium. In comparison, the ability of synthetic membrane stabilizers to protect fragile DMD skeletal muscles has been less clear. Because cardiac and skeletal muscles have distinct structural and functional features, including differences in the mechanism of activation, variance in sarcolemma phospholipid composition, and differences in the magnitude and types of forces generated, we speculated that optimized membrane stabilization could be inherently different. Our objective here is to use principles of pharmacodynamics to evaluate membrane stabilization therapy for DMD skeletal muscles. Results show a dramatic differential effect of membrane stabilization by optimization of pharmacodynamic-guided route of poloxamer delivery. Data show that subcutaneous P188 delivery, but not intravascular or intraperitoneal routes, conferred significant protection to dystrophic limb skeletal muscles undergoing mechanical stress in vivo. In addition, structure-function examination of synthetic membrane stabilizers further underscores the importance of copolymer composition, molecular weight, and dosage in optimization of poloxamer pharmacodynamics in vivo.

Conflict of interest statement

J.M.M. is on the scientific advisory board of and holds shares in Phrixus Pharmaceuticals Inc., a company developing novel therapeutics for heart failure.

Figures

Figure 1
Figure 1
In vitro hypo-osmotic stress assay to screen block copolymers for membrane stabilization. (a) Schematic representation of the tested triblock copolymers and the exclusively polyethylene oxide (PEO)-based polymer PEG8000, showing relative polypropylene oxide (PPO) (red) and PEO composition (blue). (b) Satellite cell-derived myotubes from normal muscle were exposed to hypo-osmotic stress media and lactate dehydrogenase (LDH) enzyme release into the media was measured as a marker of membrane leak. The ability of P188 and other copolymers to decrease LDH leakage from stressed myotubes was assessed as a parameter of membrane stabilization. Data are presented as % LDH release normalized to total LDH release (*P < 0.05, via one-way analysis of variance compared to nontreated group, #P < 0.05 indicates 150 µmol/l P188 is significantly different from 1 µmol/l P188, &P < 0.05 indicates 1 µmol/l ext-P188 is significantly different from 0.1 µmol/l ext-P188). Mean values are derived from at least three independent experiments with three to four wells of densely cultured myotubes per experiment. Error bars shown as mean ± standard error of the mean.
Figure 2
Figure 2
Membrane stabilizers P188 and ext-P188 decrease hypo-osmotic stress-induced release of LDH in dystrophic myofibers and myotubes in vitro. (a) Isolated mdx flexor digitorum brevis (FDB) myofibers and (b) double knockout (mdx/utr–/–) myotubes were subjected to hypo-osmotic stress and lactate dehydrogenase (LDH) release was assessed in the absence (nontreated) or presence of increasing concentrations of P188 and ext-P188. Data presented as % LDH release normalized to total LDH release from nontreated. *P < 0.05 via one-way analysis of variance compared to nontreated group. Mean values are derived from at least three independent experiments with three to four wells of 20–30 isolated mdx FDB myofibers or densely cultured mdx/utr–/– myotubes per experiment. Error bars shown as mean ± standard error of the mean.
Figure 3
Figure 3
Intraperitoneal and intravenous delivery of P188 have no effect to protect against lengthening contraction-induced force loss in mdx mice in vivo. Force loss by the anterior crural muscles in adult mdx mice (n = 5–8 per group) treated with 460 mg/kg P188 or saline vehicle either (a) intraperitoneally (IP) or (b) intravenously (IV), at least 30 minutes before the injury protocol was assessed over the course of 50 lengthening contractions. Force loss is presented as a fraction of the initial maximal force ± standard error of the mean (SEM). Contractions #1–15 show the marked force deficit in mdx mice. Results from BL/10 control mice injected saline via each delivery route are also shown. Inset (a,b) shows all 50 contractions. Both IP and IV delivery of P188 had no significant effect to decrease force loss over the course of the protocol. Error bars shown as mean ± SEM (in some cases symbol size was larger than error bar).
Figure 4
Figure 4
Subcutaneous delivery of P188 markedly protects against lengthening contraction-induced force loss in mdx mice in vivo. (a) Representative individual force tracings during lengthening contractions #1 versus #10 for subQ saline and subQ P188 treated mdx mice. (b) SubQ P188 delivery confers significant protection against force loss from contractions #5–50 (*P < 0.05 via two-way analysis of variance compared to the mdx saline group. Contractions #1–15 are highlighted with the entire protocol shown in inset. Error bars shown as mean ± SEM (in some cases symbol size was larger than error bar).
Figure 5
Figure 5
P188 significantly decreases baseline and lengthening contraction-induced sarcolemmal instability. (a) Example images of Evans Blue dye uptake into TA myofibers 2 hours postinjury in mdx mice treated subQ with saline or P188. (b) Quantification of total Evans Blue Dye extracted from whole TA muscles. Evans Blue dye was extracted by incubating the whole TA muscle in 1 ml of formamide. “C” denotes the contralateral noninjured TA, “LC” denotes lengthening contraction. Data is shown as absorbance at 620 nm divided by mg of tissue. **P < 0.01, ***P < 0.001 via one-way analysis of variance. Error bars shown as mean ± standard error of the mean.
Figure 6
Figure 6
In vivo isometric force measurements in P188-treated mdx mice immediately postinjury. Intraperitoneal (a) and intravenous (b) P188 delivery had no significant effect on peak isometric force immediately after injury whereas subQ P188 treatment (c) significantly enhanced immediate postinjury isometric force recovery compared to mdx saline, and was not significantly different from BL/10 saline group. *P < 0.05 one-way analysis of variance compared to mdx saline, #P < 0.05 one-way analysis of variance compared to BL/10 saline. Error bars shown as mean ± standard error of the mean.
Figure 7
Figure 7
Dose and delivery route dependence of ext-P188 in protecting mdx mice skeletal muscle from lengthening contraction-induced force loss in vivo. Mdx mice were administered either low dose (60 mg/kg) (gray symbols) or high dose (160 mg/kg) (black symbols) ext-P188 via intraperitoneal (a) (IP), (b) intravenous (IV)), or (c) subcutaneous (subQ) injection at least 30 minutes prior to the protocol and the anterior crural muscle group tested for lengthening contraction-induced force loss over the course of 50 contractions. Force loss is presented as a fraction of the initial maximal force ± standard error of the mean (SEM). Contractions #1–15 are graphically emphasized. (a) High-dose ext-P188 (160 mg/kg) had no protective effect against lengthening contraction-induced force loss. However, low dose ext-P188 (60 mg/kg) significantly improved resistance to injury from contraction #8 through #50 (*P < 0.05 via two-way analysis of variance compared to mdx saline). (b,c) Neither low-dose nor high-dose ext-P188 administered IV or subQ had any significant effects on force deficit. (d) Low dosage 60 mg/kg ext-P188 was injected into the tibialis anterior (TA) muscle of mdx mice intramuscularly (IM) directly prior to the protocol. Ext-P188-treated mice had a highly significant and sustained resistance to lengthening contraction injury loss (contractions 6–50, *P < 0.05 two-way analysis of variance compared to mdx saline). Contractions #1–15 are graphically emphasized, entire protocol shown in inset. Error bars shown as mean ± SEM (in some cases symbol size was larger than error bar).
Figure 8
Figure 8
Isometric force measurements in ext-P188-treated mdx immediately postinjury. (a) Low-dose (60 mg/kg) ext-P188 delivered IP significantly prevented postlengthening protocol isometric force loss (*P < 0.05 one-way analysis of variance compared to mdx saline) although was significantly different from the BL/10 saline group loss (#P < 0.05 one-way analysis of variance compared to BL/10 saline). (b,c). No significant effect obtained for IV or subQ ext-P188-treated mice. (d) Postinjury isometric force of mdx mice treated with direct injection of 60 mg/kg ext-P188 into the TA muscle was not significantly different from BL/10 saline (#P = 0.74, one-way analysis of variance compared to BL/10 saline). Error bars shown as mean ± standard error of the mean.

References

    1. Emery, A (2002). Duchenne muscular dystrophy. Motulski AG 10: 138–51.
    1. Hoffman, EP, Brown, RH Jr and Kunkel, LM (1987). Dystrophin: the protein product of the Duchenne muscular dystrophy locus. Cell 51: 919–928.
    1. Eagle, M, Bourke, J, Bullock, R, Gibson, M, Mehta, J, Giddings, D et al. (2007). Managing Duchenne muscular dystrophy–the additive effect of spinal surgery and home nocturnal ventilation in improving survival. Neuromuscul Disord 17: 470–475.
    1. Finsterer, J and Stöllberger, C (2003). The heart in human dystrophinopathies. Cardiology 99: 1–19.
    1. Moriuchi, T, Kagawa, N, Mukoyama, M and Hizawa, K (1993). Autopsy analyses of the muscular dystrophies. Tokushima J Exp Med 40: 83–93.
    1. Manzur, AY, Kuntzer, T, Pike, M and Swan, A (2008). Glucocorticoid corticosteroids for Duchenne muscular dystrophy. Cochrane Database Syst Rev, CD003725.
    1. Markham, LW, Spicer, RL, Khoury, PR, Wong, BL, Mathews, KD and Cripe, LH (2005). Steroid therapy and cardiac function in Duchenne muscular dystrophy. Pediatr Cardiol 26: 768–771.
    1. Wolthers, OD and Pedersen, S (1990). Short term linear growth in asthmatic children during treatment with prednisolone. BMJ 301: 145–148.
    1. Avioli, LV (1993). Glucocorticoid effects on statural growth. Br J Rheumatol 32 Suppl 2: 27–30.
    1. McNally, EM (2007). New approaches in the therapy of cardiomyopathy in muscular dystrophy. Annu Rev Med 58: 75–88.
    1. Goyenvalle, A, Seto, JT, Davies, KE and Chamberlain, J (2011). Therapeutic approaches to muscular dystrophy. Hum Mol Genet 20(R1): R69–R78.
    1. Lu, QL, Rabinowitz, A, Chen, YC, Yokota, T, Yin, H, Alter, J et al. (2005). Systemic delivery of antisense oligoribonucleotide restores dystrophin expression in body-wide skeletal muscles. Proc Natl Acad Sci USA 102: 198–203.
    1. Alter, J, Lou, F, Rabinowitz, A, Yin, H, Rosenfeld, J, Wilton, SD et al. (2006). Systemic delivery of morpholino oligonucleotide restores dystrophin expression bodywide and improves dystrophic pathology. Nat Med 12: 175–177.
    1. Malerba, A, Sharp, PS, Graham, IR, Arechavala-Gomeza, V, Foster, K, Muntoni, F et al. (2011). Chronic systemic therapy with low-dose morpholino oligomers ameliorates the pathology and normalizes locomotor behavior in mdx mice. Mol Ther 19: 345–354.
    1. Yokota, T, Lu, QL, Partridge, T, Kobayashi, M, Nakamura, A, Takeda, S et al. (2009). Efficacy of systemic morpholino exon-skipping in Duchenne dystrophy dogs. Ann Neurol 65: 667–676.
    1. Hoffman, EP and McNally, EM (2014). Exon-skipping therapy: a roadblock, detour, or bump in the road? Sci Transl Med 6: 230fs14.
    1. Lu, QL, Cirak, S and Partridge, T (2014). What Can We Learn From Clinical Trials of Exon Skipping for DMD? Mol Ther Nucleic Acids 3: e152.
    1. Townsend, D, Yasuda, S and Metzger, J (2007). Cardiomyopathy of Duchenne muscular dystrophy: pathogenesis and prospect of membrane sealants as a new therapeutic approach. Expert Rev Cardiovasc Ther 5: 99–109.
    1. Townsend, D, Yasuda, S, Chamberlain, J and Metzger, JM (2009). Cardiac consequences to skeletal muscle-centric therapeutics for Duchenne muscular dystrophy. Trends Cardiovasc Med 19: 50–55.
    1. Petrof, BJ, Shrager, JB, Stedman, HH, Kelly, AM and Sweeney, HL (1993). Dystrophin protects the sarcolemma from stresses developed during muscle contraction. Proc Natl Acad Sci USA 90: 3710–3714.
    1. Danialou, G, Comtois, AS, Dudley, R, Karpati, G, Vincent, G, Des Rosiers, C et al. (2001). Dystrophin-deficient cardiomyocytes are abnormally vulnerable to mechanical stress-induced contractile failure and injury. FASEB J 15: 1655–1657.
    1. Yasuda, S, Townsend, D, Michele, DE, Favre, EG, Day, SM and Metzger, JM (2005). Dystrophic heart failure blocked by membrane sealant poloxamer. Nature 436: 1025–1029.
    1. Townsend, D, Turner, I, Yasuda, S, Martindale, J, Davis, J, Shillingford, M et al. (2010). Chronic administration of membrane sealant prevents severe cardiac injury and ventricular dilatation in dystrophic dogs. J Clin Invest 120: 1140–1150.
    1. Spurney, CF, Guerron, AD, Yu, Q, Sali, A, van der Meulen, JH, Hoffman, EP et al. (2011). Membrane sealant Poloxamer P188 protects against isoproterenol induced cardiomyopathy in dystrophin deficient mice. BMC Cardiovasc Disord 11: 20.
    1. Markham, BE, Kernodle, S, Nemzek, J, Wilkinson, JE and Sigler, R (2015). Chronic Dosing with Membrane Sealant Poloxamer 188 NF Improves Respiratory Dysfunction in Dystrophic Mdx and Mdx/Utrophin-/- Mice. PLoS One 10: e0134832.
    1. Quinlan, JG, Wong, BL, Niemeier, RT, McCullough, AS, Levin, L and Emanuele, M (2006). Poloxamer 188 failed to prevent exercise-induced membrane breakdown in mdx skeletal muscle fibers. Neuromuscul Disord 16: 855–864.
    1. Terry, RL, Kaneb, HM and Wells, DJ (2014). Poloxamer [corrected] 188 has a deleterious effect on dystrophic skeletal muscle function. PLoS One 9: e91221.
    1. Lee, RC, River, LP, Pan, FS, Ji, L and Wollmann, RL (1992). Surfactant-induced sealing of electropermeabilized skeletal muscle membranes in vivo. Proc Natl Acad Sci USA 89: 4524–4528.
    1. Murphy, AD, McCormack, MC, Bichara, DA, Nguyen, JT, Randolph, MA, Watkins, MT et al. (2010). Poloxamer 188 protects against ischemia-reperfusion injury in a murine hind-limb model. Plast Reconstr Surg 125: 1651–1660.
    1. Suzuki, N, Akiyama, T, Takahashi, T, Komuro, H, Warita, H, Tateyama, M et al. (2012). Continuous administration of poloxamer 188 reduces overload-induced muscular atrophy in dysferlin-deficient SJL mice. Neurosci Res 72: 181–186.
    1. Ng, R, Metzger, JM, Claflin, DR and Faulkner, JA (2008). Poloxamer 188 reduces the contraction-induced force decline in lumbrical muscles from mdx mice. Am J Physiol Cell Physiol 295: C146–C150.
    1. Wang, JY, Chin, J, Marks, JD and Lee, KY (2010). Effects of PEO-PPO-PEO triblock copolymers on phospholipid membrane integrity under osmotic stress. Langmuir 26: 12953–12961.
    1. Maskarinec, SA and Lee, KYC (2003). Comparative study of poloxamer insertion into lipid monolayers. Langmuir 19: 1809–1815.
    1. Wang, JY, Marks, J and Lee, KY (2012). Nature of interactions between PEO-PPO-PEO triblock copolymers and lipid membranes: (I) effect of polymer hydrophobicity on its ability to protect liposomes from peroxidation. Biomacromolecules 13: 2616–2623.
    1. Abdel-Rahman, SM and Kauffman, RE (2004). The integration of pharmacokinetics and pharmacodynamics: understanding dose-response. Annu Rev Pharmacol Toxicol 44: 111–136.
    1. Rowland, M and Tozer, TN. Clinical pharmacokinetics and pharmacodynamics concepts and applications., 4th ed. Philadephia (PA): Lippincott, Williams & Wilkins.
    1. Ingalls, CP, Warren, GL, Lowe, DA, Boorstein, DB and Armstrong, RB (1996). Differential effects of anesthetics on in vivo skeletal muscle contractile function in the mouse. J Appl Physiol (1985) 80: 332–340.
    1. Lovering, RM and De Deyne, PG (2004). Contractile function, sarcolemma integrity, and the loss of dystrophin after skeletal muscle eccentric contraction-induced injury. Am J Physiol Cell Physiol 286: C230–C238.
    1. Lovering, RM, Roche, JA, Bloch, RJ and De Deyne, PG (2007). Recovery of function in skeletal muscle following 2 different contraction-induced injuries. Arch Phys Med Rehabil 88: 617–625.
    1. Baltgalvis, KA, Call, JA, Nikas, JB and Lowe, DA (2009). Effects of prednisolone on skeletal muscle contractility in mdx mice. Muscle Nerve 40: 443–454.
    1. Call, JA, Warren, GL, Verma, M and Lowe, DA (2013). Acute failure of action potential conduction in mdx muscle reveals new mechanism of contraction-induced force loss. J Physiol 591(Pt 15): 3765–3776.
    1. Menke, A and Jockusch, H (1991). Decreased osmotic stability of dystrophin-less muscle cells from the mdx mouse. Nature 349: 69–71.
    1. Menke, A and Jockusch, H (1995). Extent of shock-induced membrane leakage in human and mouse myotubes depends on dystrophin. J Cell Sci 108 (Pt 2): 727–733.
    1. Cheng, CY, Wang, JY, Kausik, R, Lee, KY and Han, S (2012). Nature of interactions between PEO-PPO-PEO triblock copolymers and lipid membranes: (II) role of hydration dynamics revealed by dynamic nuclear polarization. Biomacromolecules 13: 2624–2633.
    1. Liu, M, Yue, Y, Harper, SQ, Grange, RW, Chamberlain, JS and Duan, D (2005). Adeno-associated virus-mediated microdystrophin expression protects young mdx muscle from contraction-induced injury. Mol Ther 11: 245–256.
    1. Yue, Y, Liu, M and Duan, D (2006). C-terminal-truncated microdystrophin recruits dystrobrevin and syntrophin to the dystrophin-associated glycoprotein complex and reduces muscular dystrophy in symptomatic utrophin/dystrophin double-knockout mice. Mol Ther 14: 79–87.
    1. Foster, H, Sharp, PS, Athanasopoulos, T, Trollet, C, Graham, IR, Foster, K et al. (2008). Codon and mRNA sequence optimization of microdystrophin transgenes improves expression and physiological outcome in dystrophic mdx mice following AAV2/8 gene transfer. Mol Ther 16: 1825–1832.
    1. Call, JA, Eckhoff, MD, Baltgalvis, KA, Warren, GL and Lowe, DA (2011). Adaptive strength gains in dystrophic muscle exposed to repeated bouts of eccentric contraction. J Appl Physiol (1985) 111: 1768–1777.
    1. Goudenege, S, Lamarre, Y, Dumont, N, Rousseau, J, Frenette, J, Skuk, D et al. (2010). Laminin-111: a potential therapeutic agent for Duchenne muscular dystrophy. Mol Ther 18: 2155–2163.
    1. Belanto, JJ, Mader, TL, Eckhoff, MD, Strandjord, DM, Banks, GB, Gardner, MK et al. (2014). Microtubule binding distinguishes dystrophin from utrophin. Proc Natl Acad Sci USA 111: 5723–5728.
    1. Roche, JA, Lovering, RM and Bloch, RJ (2008). Impaired recovery of dysferlin-null skeletal muscle after contraction-induced injury in vivo. Neuroreport 19: 1579–1584.
    1. Verma, M, Asakura, Y, Hirai, H, Watanabe, S, Tastad, C, Fong, GH et al. (2010). Flt-1 haploinsufficiency ameliorates muscular dystrophy phenotype by developmentally increased vasculature in mdx mice. Hum Mol Genet 19: 4145–4159.
    1. Pratt, SJ, Shah, SB, Ward, CW, Inacio, MP, Stains, JP and Lovering, RM (2013). Effects of in vivo injury on the neuromuscular junction in healthy and dystrophic muscles. J Physiol 591(Pt 2): 559–570.
    1. Arpke, RW, Darabi, R, Mader, TL, Zhang, Y, Toyama, A, Lonetree, CL et al. (2013). A new immuno-, dystrophin-deficient model, the NSG-mdx(4Cv) mouse, provides evidence for functional improvement following allogeneic satellite cell transplantation. Stem Cells 31: 1611–1620.
    1. Harper, SQ, Hauser, MA, DelloRusso, C, Duan, D, Crawford, RW, Phelps, SF et al. (2002). Modular flexibility of dystrophin: implications for gene therapy of Duchenne muscular dystrophy. Nat Med 8: 253–261.
    1. Goyenvalle, A, Griffith, G, Babbs, A, El Andaloussi, S, Ezzat, K, Avril, A et al. (2015). Functional correction in mouse models of muscular dystrophy using exon-skipping tricyclo-DNA oligomers. Nat Med 21: 270–275.
    1. Turner, PV, Brabb, T, Pekow, C and Vasbinder, MA (2011). Administration of substances to laboratory animals: routes of administration and factors to consider. J Am Assoc Lab Anim Sci 50: 600–613.
    1. McLennan, DN, Porter, CJ and Charman, SA (2005). Subcutaneous drug delivery and the role of the lymphatics. Drug Discov Today Technol 2: 89–96.
    1. Holland, RJ, Parker, EJ, Guiney, K and Zeld, FR (1995). Fluorescence probe studies of ethylene oxide/propylene oxide block copolymers in aqueous solution. J Phys Chem 99: 11981–11988.
    1. Kabanov, AV, Batrakova, EV and Alakhov, VY (2002). Pluronic block copolymers as novel polymer therapeutics for drug and gene delivery. J Control Release 82: 189–212.
    1. Kabanov, AV, Batrakova, EV and Miller, DW (2003). Pluronic block copolymers as modulators of drug efflux transporter activity in the blood-brain barrier. Adv Drug Deliv Rev 55: 151–164.
    1. Alexandridis, P and Hatton, TA (1995). Poly(ethylene oxide)-poly(propylene oxide)-poly(ethylene oxide) block-copolymer surfactants in aqueous-solutions and at interfaces - thermodynamics, structure, dynamics, and modeling. Colloids Surfaces A-Physicochemical Eng Asp 96: 1–46.
    1. Bers, DM (2002). Cardiac excitation-contraction coupling. Nature 415: 198–205.
    1. de Kretser, TA and Livett, BG (1977). Skeletal-muscle sarcolemma from normal and dystrophic mice. Isolation, characterization and lipid composition. Biochem J 168: 229–237.
    1. Otten, W, Iaizzo, PA and Eichinger, HM (1997). Effects of a high n-3 fatty acid diet on membrane lipid composition of heart and skeletal muscle in normal swine and in swine with the genetic mutation for malignant hyperthermia. J Lipid Res 38: 2023–2034.
    1. Thomas, GD, Sander, M, Lau, KS, Huang, PL, Stull, JT and Victor, RG (1998). Impaired metabolic modulation of α-adrenergic vasoconstriction in dystrophin-deficient skeletal muscle. Proc Natl Acad Sci USA 95: 15090–15095.
    1. Bagher, P, Duan, D and Segal, SS (2011). Evidence for impaired neurovascular transmission in a murine model of Duchenne muscular dystrophy. J Appl Physiol (1985) 110: 601–609.
    1. Frey, SL, Zhang, D, Carignano, MA, Szleifer, I and Lee, KY (2007). Effects of block copolymer’s architecture on its association with lipid membranes: experiments and simulations. J Chem Phys 127: 114904.
    1. Leung, DG, Herzka, DA, Thompson, WR, He, B, Bibat, G, Tennekoon, G et al. (2014). Sildenafil does not improve cardiomyopathy in Duchenne/Becker muscular dystrophy. Ann Neurol 76: 541–549.
    1. Mendell, JR, Campbell, K, Rodino-Klapac, L, Sahenk, Z, Shilling, C, Lewis, S et al. (2010). Dystrophin immunity in Duchenne’s muscular dystrophy. N Engl J Med 363: 1429–1437.
    1. Kabanov, AV, Batrakova, EV, Sriadibhatla, S, Yang, Z, Kelly, DL and Alakov, VY (2005). Polymer genomics: shifting the gene and drug delivery paradigms. J Control Release 101: 259–271.
    1. Jewell, RC, Khor, SP, Kisor, DF, LaCroix, KA and Wargin, WA (1997). Pharmacokinetics of RheothRx injection in healthy male volunteers. J Pharm Sci 86: 808–812.
    1. Ballas, SK, Files, B, Luchtman-Jones, L, Benjamin, L, Swerdlow, P, Hilliard, L et al. (2004). Safety of purified poloxamer 188 in sickle cell disease: phase I study of a non-ionic surfactant in the management of acute chest syndrome. Hemoglobin 28: 85–102.
    1. Martindale, JJ and Metzger, JM (2014). Uncoupling of increased cellular oxidative stress and myocardial ischemia reperfusion injury by directed sarcolemma stabilization. J Mol Cell Cardiol 67: 26–37.
    1. Bosnakovski, D, Xu, Z, Li, W, Thet, S, Cleaver, O, Perlingeiro, RC et al. (2008). Prospective isolation of skeletal muscle stem cells with a Pax7 reporter. Stem Cells 26: 3194–3204.
    1. Filareto, A, Parker, S, Darabi, R, Borges, L, Iacovino, M, Schaaf, T et al. (2013). An ex vivo gene therapy approach to treat muscular dystrophy using inducible pluripotent stem cells. Nat Commun 4: 1549.
    1. Darabi, R, Gehlbach, K, Bachoo, RM, Kamath, S, Osawa, M, Kamm, KE et al. (2008). Functional skeletal muscle regeneration from differentiating embryonic stem cells. Nat Med 14: 134–143.
    1. Matsuda, R, Nishikawa, A and Tanaka, H (1995). Visualization of dystrophic muscle fibers in mdx mouse by vital staining with Evans blue: evidence of apoptosis in dystrophin-deficient muscle. J Biochem 118: 959–964.
    1. Heydemann, A, Huber, JM, Demonbreun, A, Hadhazy, M and McNally, EM (2005). Genetic background influences muscular dystrophy. Neuromuscul Disord 15: 601–609.

Source: PubMed

3
Abonneren