Poloxamer 188 enhances functional recovery of lethally heat-shocked fibroblasts.

Damage to the cell membrane has been implicated as the primary event in the pathogenesis of heat shock, generally resulting in loss of cellular homeostasis and cell death. Thus a promising mode of therapy would involve the restoration of cell membrane integrity. Surfactant molecules, specifically triblock polymers such as Poloxamer 188 (P-188), possess the ability to self-aggregate into membrane-like structures in aqueous solutions and have been shown to restore membrane integrity. The objective of this study was to develop functional and morphological assays to determine whether treatment with P-188 after heat shock enhances the recovery of thermally damaged cells. Human foreskin fibroblasts were placed in sterile vials and heated by immersion in a calibrated water bath for various lengths of time at predefined temperatures. Cell recovery after heat shock was assessed using a functional assay based on the ability of the cells to contract fibroblast populated collagen lattices (FPCLs). Subsequent to heating, collagen lattices were prepared with control (no heat, no P-188) and heat shocked cells (with and without P-188). Our results indicate that treatment with low concentrations of P-188 after heat shock was effective in ameliorating both the morphological integrity and the contractile function of thermally damaged cells. Further, we observed that P-188 was most effective in improving the contractile ability of cells heat shocked at 45 degrees C; however, it had no influence on the contractility of cells exposed to higher temperatures. Our results suggest that there exists a threshold of thermal stress (45 degrees C for 20-60 min) beyond which treatment with low concentrations of P-188 (0.5 mg/ml) is ineffective in minimizing cell damage. Moreover, the results of our morphological assays indicate that cells treated with P-188 after heat shock maintain their cytoskeletal organization, whereas untreated cells exhibit filamentous actin depolymerization.

[1]  W. Dewey,et al.  A direct correlation between hyperthermia‐induced membrane blebbing and survival in synchronous G1 CHO cells , 1986, Journal of cellular physiology.

[2]  L Tung,et al.  Poloxamer 188 decreases susceptibility of artificial lipid membranes to electroporation. , 1996, Biophysical journal.

[3]  R. Hotchkiss,et al.  Mechanisms of cell injury and death. , 1996, British journal of anaesthesia.

[4]  J C Bischof,et al.  Dynamics of cell membrane permeability changes at supraphysiological temperatures. , 1995, Biophysical journal.

[5]  F. Grinnell,et al.  Reorganization of hydrated collagen lattices by human skin fibroblasts. , 1984, Journal of cell science.

[6]  J. Lepock,et al.  Involvement of membranes in cellular responses to hyperthermia. , 1982, Radiation research.

[7]  W. Dewey,et al.  Failla memorial lecture. The search for critical cellular targets damaged by heat. , 1989, Radiation research.

[8]  P. Lin,et al.  Hyperthermic treatment (43 degrees C) rapidly impedes attachment of fibroblasts to culture substrates. , 1977, Cell biology international reports.

[9]  J. Lévêque,et al.  Measurement of mechanical forces generated by skin fibroblasts embedded in a three-dimensional collagen gel. , 1991, The Journal of investigative dermatology.

[10]  I. R. Schmolka Physical Basis for Poloxamer Interactions , 1994, Annals of the New York Academy of Sciences.

[11]  R. Benz,et al.  Use of irreversible electrical breakdown of lipid bilayers for the study of interaction of membranes with surface active molecules. , 1993, Biochimica et biophysica acta.

[12]  S. Benita,et al.  Interactions of a non-ionic ABA copolymer surfactant with phospholipid monolayers: Possible relevance to emulsion stabilization , 1991 .

[13]  J. Crapo,et al.  Biology of disease: free radicals and tissue injury. , 1982, Laboratory investigation; a journal of technical methods and pathology.

[14]  R. Winward,et al.  Factors affecting the heat-induced increase in protein content of chromatin. , 1980, Radiation research.

[15]  W. Dewey,et al.  Evaluation of a role for intracellular Na+, K+, Ca2+, and Mg2+ in hyperthermic cell killing. , 1986, Radiation research.

[16]  H. Meiselman,et al.  Inhibition of red blood cell-induced platelet aggregation in whole blood by a nonionic surfactant, poloxamer 188 (RheothRx injection). , 1995, Thrombosis research.

[17]  M. Toner,et al.  Effectiveness of Poloxamer 188 in Arresting Calcein Leakage from Thermally Damaged Isolated Skeletal Muscle Cells a , 1994, Annals of the New York Academy of Sciences.

[18]  R. Gibbons,et al.  Beneficial effects of RheothRx injection in patients receiving thrombolytic therapy for acute myocardial infarction. Results of a randomized, double-blind, placebo-controlled trial. , 1996, Circulation.

[19]  H. Meiselman,et al.  Haemorheological effects of a nonionic copolymer surfactant (Poloxamer 188) , 1992 .

[20]  R. Lee,et al.  Surfactant-induced sealing of electropermeabilized skeletal muscle membranes in vivo. , 1992, Proceedings of the National Academy of Sciences of the United States of America.

[21]  J. Parrillo,et al.  Reduction in Reperfusion‐Induced Myocardial Necrosis in Dogs by RheothRx Injection (Poloxamer 188 N.F.), a Hemorheological Agent That Alters Neutrophil Function , 1994, Circulation.

[22]  B. Magun,et al.  Effects of hyperthermia on binding, internalization, and degradation of epidermal growth factor. , 1981, Radiation research.

[23]  Raphael C. Lee,et al.  Promising Therapy for Cell Membrane Damage , 1994, Annals of the New York Academy of Sciences.

[24]  P S Lin,et al.  Temperature-induced variations in the surface topology of cultured lymphocytes are revealed by scanning electron microscopy. , 1973, Proceedings of the National Academy of Sciences of the United States of America.

[25]  W. Dewey,et al.  Effects of hyperthermia on dividing Chinese hamster ovary cells and on microtubules in vitro. , 1982, Cancer research.

[26]  J. Teissié,et al.  Spatial compartmentation and time resolution of photooxidation of a cell membrane probe in electropermeabilized Chinese hamster ovary cells. , 1995, European journal of biochemistry.

[27]  M. Steinberg,et al.  RheothRx (poloxamer 188) injection for the acute painful episode of sickle cell disease: a pilot study. , 1997, Blood.

[28]  H. Ehrlich,et al.  Differences in cell division and thymidine incorporation with rat and primate fibroblasts in collagen lattices. , 1992, Tissue & cell.

[29]  H. Ehrlich,et al.  Physiological variables affecting collagen lattice contraction by human dermal fibroblasts. , 1989, Experimental and molecular pathology.

[30]  C. Heidelberger,et al.  Studies on the quantitative biology of hyperthermic killing of HeLa cells. , 1973, Cancer research.

[31]  E Bell,et al.  Production of a tissue-like structure by contraction of collagen lattices by human fibroblasts of different proliferative potential in vitro. , 1979, Proceedings of the National Academy of Sciences of the United States of America.

[32]  Y. Wu,et al.  Changes in membrane properties during energy depletion-induced cell injury studied with fluorescence microscopy. , 1996, Biophysical journal.

[33]  M. Capelli-Schellpfeffer,et al.  Advances in the evaluation and treatment of electrical and thermal injury emergencies , 1994, Proceedings of IEEE Petroleum and Chemical Industry Technical Conference (PCIC '94).

[34]  Ernest G. Cravalho,et al.  Time Progression of Hemolysis of Erythrocyte Populations Exposed to Supraphysiological Temperatures , 1979 .

[35]  G. Bartosz,et al.  Membrane effects of ionizing radiation and hyperthermia. , 1986, International journal of radiation biology and related studies in physics, chemistry, and medicine.

[36]  C. Sato,et al.  Effects of hyperthermia on cell surface charge and cell survival in mastocytoma cells. , 1981, Cancer research.