:::

詳目顯示

回上一頁
題名:不同阻力與加壓負荷對合成激素與代謝壓力之影響
作者:周峻忠 引用關係
作者(外文):Chun-Chung Chou
校院名稱:中國文化大學
系所名稱:體育學系運動教練碩博士班
指導教授:林正常
學位類別:博士
出版日期:2011
主題關鍵詞:加壓運動生長激素睪固酮肌電圖occlusion exercisegrowth hormonetestosteroneelectromyography
原始連結:連回原系統網址new window
相關次數:
  • 被引用次數被引用次數:期刊(0) 博士論文(0) 專書(0) 專書論文(0)
  • 排除自我引用排除自我引用:0
  • 共同引用共同引用:0
  • 點閱點閱:51
目的:探討不同阻力(70% 1RM 與 40% 1RM)搭配不同加壓負荷(1.3倍收縮壓壓力與0.7倍的收縮壓壓力)對運動後立即的合成激素反應與代謝壓力的影響。方法:招募15名一般健康男性(年齡:19.8 ± 1.6歲,身高:171.1 ± 4.5 公分,體重:67.3 ± 7.8公斤),依平衡次序原則,讓研究對象分別接受下列五種實驗處理:(一)高強度阻力運動 (HR, 70% 1RM);(二)低強度阻力運動 (LR, 40% 1RM);(三)高強度阻力運動 + 低加壓測驗 (HRLO, 70% 1RM + 0.7 SBP);(四)低強度阻力運動 + 高加壓測驗 (LRHO, 40% 1RM + 1.3 SBP)與(五)低強度阻力運動 + 低加壓測驗 (LRLO, 40% 1RM + 0.7 SBP),實驗處理間休息至少五日。阻力運動以斜坐推蹬方式進行雙腿阻力運動(共5組,每組反覆12次,組間休息1分鐘);加壓部位於雙腿大腿近端(加壓時間與放鬆時間同阻力運動)。在運動前 (pre)、運動後立即 (post)、15 (post-15)、30 (post-30)、60 (post-60) 分鐘進行抽血,比較血清胰島素、血清睪固酮、血清生長激素、血清皮質固醇、血漿乳酸、pH與血漿葡萄糖之差異;並比較實驗處理前 (pre) 後 (post-60) 對最大肌力(最大等長肌力、最大等張肌力)與肌電訊號(股外側肌)之差異。結果:HRLO在運動後有較高的生長激素且持續至運動後30分鐘 (pre: 0.95 ± 2.07, post: 6.01 ± 6.24, post-15: 4.53 ± 4.51, post-30: 3.44 ±3.23 ng / ml, p > .05),HRLO相較於LRHO、LRLO與LR在運動後立即 (HRLO: 6.80 ± 1.77, LRHO: 5.82 ± 1.53, LRLO: 5.47 ± 1.18, LR: 5.74 ± 1.48, p < .05) 與運動後60分鐘 (HRLO: 6.35 ± 1.53, LRHO: 5.19 ± 1.17, LRLO: 5.32 ± 1.08, LR: 5.54 ± 1.47, p < .05) 有較高的睪固酮濃度;HRLO的實驗處理能在運動後引起較高的代謝壓力反應 (LA、pH值與RPE) 並降低最大等長肌力表現(下降9.3%)。結論:高阻力低加壓的運動模式引起體內較高的合成激素反應,雖產生較大的代謝壓力反應伴隨肌力下降的現象,但此一立即性的反應是造成細胞蛋白質合成與肌肥大之關鍵基礎。
Purpose: The aim of this study was to investigate the effect of different resistance loads (70% 1RM and 40% 1RM) combined different occlusion pressures (1.3SBP and 0.7SBP) on post-exercise acute anabolic hormone and metabolic stress. Methods: Fifteen healthy males (age: 19.8 ± 1.6yrs; height: 171.1 ± 4.5 cm; weight: 67.3 ± 7.8kg) were recruited in this study to perform bilateral leg extension (5 sets, 12 repetitions with 1min rest among all sets) under five experimental conditions: (1) high intensity resistance exercise (HR, 70% 1RM); (2) low intensity resistance exercise (LR, 40% 1RM); (3) high intensity resistance exercise with low occlusion pressure (HRLO, 70% 1RM + 0.7SBP); (4) low intensity resistance exercise with high occlusion pressure (LRHO, 40% 1RM + 1.3SBP); (5) low intensity resistance exercise with low occlusion pressure (LRLO, 40% 1RM + 0.7SBP). Blood samples were collected prior exercise (pre), and 0, 15, 30, 60 min after exercise (post, post-15, post-30 and post-60) for analysis of growth hormone, testosterone, cortisol, insulin, glucose, lactic acid and pH. The maximal voluntary contraction and electromyography of vastus lateralis were compared at pre and post-60. By counter balance design, each experiment separated by at least 5 days. Results: The GH of HRLO was higher at post, post-15 and post-30 than pre (pre: 0.95 ± 2.07, post: 6.01 ± 6.24, post-15: 4.53 ± 4.51, post-30: 3.44 ±3.23 ng / ml, p > .05). The testosterone of HRLO was higher at post (HRLO: 6.80 ± 1.77, LRHO: 5.82 ± 1.53, LRLO: 5.47 ± 1.18, LR: 5.74 ± 1.48, p < .05) and post-60 (HRLO: 6.35 ± 1.53, LRHO: 5.19 ± 1.17, LRLO: 5.32 ± 1.08, LR: 5.54 ± 1.47, p < .05) than LRHO, LRLO and LR. HRLO also induced higher metabolic stress (LA and pH) and more muscular strength decline (-9.3%). Conclusion: The mode of high intensity resistance loads combined low occlusion pressure could induce higher anabolic hormone response, although higher metabolic stress and more muscular strength decline were also found, but this acute response was the essential fundamental to activate muscle protein synthesis and hypertrophy.
林正常、蔡崇濱、林信甫、林政東、吳博翰、鄭景峰等(譯) (2004)。肌力與體能訓練(原作者:T. R. Baechle & R. W. Earle)。臺北市:藝軒(原著出版年:2000)。
高藤曉子、賴麗雲 (2006)。加壓肌力訓練法。國民體育季刊,35 (3),65-72。
廖翊宏、陳宗與、林信甫、周峻忠 (2010)。急性阻力運動促進肌肉蛋白質合成之生理與分子機制之探討。運動生理暨體能學報,10,13-28。new window
Aagaard, P., Simonsen, E. B., Andersen, J. L., Magnusson, S. P., Bojsen-Moller, F., & Dyhre-Poulsen, P. (2000). Antagonist muscle coactivation during isokinetic knee extension. Scandinavian Journal of Medicine and Science in Sports, 10(2), 58-67.
Abe, T., Kearns, C. F., & Sato, Y. (2006). Muscle size and strength are increased following walk training with restricted venous blood flow from the leg muscle, Kaatsu-walk training. Journal of Applied Physiology, 100(5), 1460-1466.
Adams, G. R. (2002). Autocrine and/or paracrine insulin-like growth factor-I activity in skeletal muscle. Clinical Orthopaedics and Related Research(403 Suppl), S188-196.
Ahtiainen, J. P., Pakarinen, A., Kraemer, W. J., & Hakkinen, K. (2003). Acute hormonal and neuromuscular responses and recovery to forced vs maximum repetitions multiple resistance exercises. International Journal of Sports Medicine, 24(6), 410-418.
Atherton, P. J., & Rennie, M. J. (2006). Protein synthesis a low priority for exercising muscle. Journal of Physiology, 573(Pt 2), 288-289.
Bamman, M. M., Shipp, J. R., Jiang, J., Gower, B. A., Hunter, G. R., Goodman, A., et al. (2001). Mechanical load increases muscle IGF-I and androgen receptor mRNA concentrations in humans. American Journal of Physiology. Endocrinology and Metabolism, 280(3), E383-390.
Banz, W. J., Maher, M. A., Thompson, W. G., Bassett, D. R., Moore, W., Ashraf, M., et al. (2003). Effects of resistance versus aerobic training on coronary artery disease risk factors. Experimental Biology and Medicine (Maywood), 228(4), 434-440.
Beltman, J. G., de Haan, A., Haan, H., Gerrits, H. L., van Mechelen, W., & Sargeant, A. J. (2004). Metabolically assessed muscle fibre recruitment in brief isometric contractions at different intensities. European Journal of Applied Physiology and Occupational Physiology, 92(4-5), 485-492.
Bolster, D. R., Kubica, N., Crozier, S. J., Williamson, D. L., Farrell, P. A., Kimball, S. R., et al. (2003). Immediate response of mammalian target of rapamycin (mTOR)-mediated signalling following acute resistance exercise in rat skeletal muscle. Journal of Physiology, 553(Pt 1), 213-220.new window
Brahm, H., Piehl-Aulin, K., Saltin, B., & Ljunghall, S. (1997). Net fluxes over working thigh of hormones, growth factors and biomarkers of bone metabolism during short lasting dynamic exercise. Calcified Tissue International, 60(2), 175-180.
Burgomaster, K. A., Moore, D. R., Schofield, L. M., Phillips, S. M., Sale, D. G., & Gibala, M. J. (2003). Resistance training with vascular occlusion: metabolic adaptations in human muscle. Medicine and Science in Sports and Exercise, 35(7), 1203-1208.
Campos, G. E., Luecke, T. J., Wendeln, H. K., Toma, K., Hagerman, F. C., Murray, T. F., et al. (2002). Muscular adaptations in response to three different resistance-training regimens: specificity of repetition maximum training zones. European Journal of Applied Physiology and Occupational Physiology, 88(1-2), 50-60.
Carter-Su, C., & Smit, L. S. (1998). Signaling via JAK tyrosine kinases: growth hormone receptor as a model system. Recent Progress in Hormone Research, 53, 61-82; discussion 82-63.
Chandler, R. M., Byrne, H. K., Patterson, J. G., & Ivy, J. L. (1994). Dietary supplements affect the anabolic hormones after weight-training exercise. Journal of Applied Physiology, 76(2), 839-845.
Chesley, A., MacDougall, J. D., Tarnopolsky, M. A., Atkinson, S. A., & Smith, K. (1992). Changes in human muscle protein synthesis after resistance exercise. Journal of Applied Physiology, 73(4), 1383-1388.
Clark, B. C., Fernhall, B., & Ploutz-Snyder, L. L. (2006). Adaptations in human neuromuscular function following prolonged unweighting: I. Skeletal muscle contractile properties and applied ischemia efficacy. Journal of Applied Physiology, 101(1), 256-263.new window
Cook, S. B., Brown, K. A., Deruisseau, K., Kanaley, J. A., & Ploutz-Snyder, L. L. (2010). Skeletal muscle adaptations following blood flow-restricted training during 30 days of muscular unloading. Journal of Applied Physiology, 109(2), 341-349.
Cook, S. B., Clark, B. C., & Ploutz-Snyder, L. L. (2007). Effects of exercise load and blood-flow restriction on skeletal muscle function. Medicine and Science in Sports and Exercise, 39(10), 1708-1713.
Copeland, K. C., Colletti, R. B., Devlin, J. T., & McAuliffe, T. L. (1990). The relationship between insulin-like growth factor-I, adiposity, and aging. Metabolism: Clinical and Experimental, 39(6), 584-587.
Cumming, D. C., Wall, S. R., Quinney, H. A., & Belcastro, A. N. (1987). Decrease in serum testosterone levels with maximal intensity swimming exercise in trained male and female swimmers. Endocrine Research, 13(1), 31-41.new window
Dohm, G. L., Tapscott, E. B., & Kasperek, G. J. (1987). Protein degradation during endurance exercise and recovery. Medicine and Science in Sports and Exercise, 19(5 Suppl), S166-171.
Dreyer, H. C., Fujita, S., Cadenas, J. G., Chinkes, D. L., Volpi, E., & Rasmussen, B. B. (2006). Resistance exercise increases AMPK activity and reduces 4E-BP1 phosphorylation and protein synthesis in human skeletal muscle. Journal of Physiology, 576(Pt 2), 613-624.
Durieux, A. C., Desplanches, D., Freyssenet, D., & Fluck, M. (2007). Mechanotransduction in striated muscle via focal adhesion kinase. Biochemical Society Transactions, 35(Pt 5), 1312-1313.
Feldman, H. A., Longcope, C., Derby, C. A., Johannes, C. B., Araujo, A. B., Coviello, A. D., et al. (2002). Age trends in the level of serum testosterone and other hormones in middle-aged men: longitudinal results from the Massachusetts male aging study. Journal of Clinical Endocrinology and Metabolism, 87(2), 589-598.
Fingar, D. C., Salama, S., Tsou, C., Harlow, E., & Blenis, J. (2002). Mammalian cell size is controlled by mTOR and its downstream targets S6K1 and 4EBP1/eIF4E. Genes and Development, 16(12), 1472-1487.
Fitts, R. H. (1994). Cellular mechanisms of muscle fatigue. Physiological Reviews, 74(1), 49-94.new window
Fujita, S., Abe, T., Drummond, M. J., Cadenas, J. G., Dreyer, H. C., Sato, Y., et al. (2007). Blood flow restriction during low-intensity resistance exercise increases S6K1 phosphorylation and muscle protein synthesis. Journal of Applied Physiology, 103(3), 903-910.
Gan, B., Yoo, Y., & Guan, J. L. (2006). Association of focal adhesion kinase with tuberous sclerosis complex 2 in the regulation of s6 kinase activation and cell growth. Journal of Biological Chemistry, 281(49), 37321-37329.
Goldspink, G. (1999). Changes in muscle mass and phenotype and the expression of autocrine and systemic growth factors by muscle in response to stretch and overload. Journal of Anatomy, 194 ( Pt 3), 323-334.
Goldspink, G. (2006). Impairment of IGF-I gene splicing and MGF expression associated with muscle wasting. International Journal of Biochemistry and Cell Biology, 38(3), 481-489.
Goto, K., Ishii, N., Kizuka, T., & Takamatsu, K. (2005). The impact of metabolic stress on hormonal responses and muscular adaptations. Medicine and Science in Sports and Exercise, 37(6), 955-963.
Gotshalk, L. A., Loebel, C. C., Nindl, B. C., Putukian, M., Sebastianelli, W. J., Newton, R. U., et al. (1997). Hormonal responses of multiset versus single-set heavy-resistance exercise protocols. Canadian Journal of Applied Physiology, 22(3), 244-255.
Hakkinen, K. (1989). Neuromuscular and hormonal adaptations during strength and power training. A review. Journal of Sports Medicine and Physical Fitness, 29(1), 9-26.new window
Hakkinen, K., Kallinen, M., Izquierdo, M., Jokelainen, K., Lassila, H., Malkia, E., et al. (1998). Changes in agonist-antagonist EMG, muscle CSA, and force during strength training in middle-aged and older people. Journal of Applied Physiology, 84(4), 1341-1349.
Hakkinen, K., & Pakarinen, A. (1993a). Acute hormonal responses to two different fatiguing heavy-resistance protocols in male athletes. Journal of Applied Physiology, 74(2), 882-887.
Hakkinen, K., & Pakarinen, A. (1993b). Muscle strength and serum testosterone, cortisol and SHBG concentrations in middle-aged and elderly men and women. Acta Physiologica Scandinavica, 148(2), 199-207.
Hakkinen, K., Pakarinen, A., Alen, M., Kauhanen, H., & Komi, P. V. (1987). Relationships between training volume, physical performance capacity, and serum hormone concentrations during prolonged training in elite weight lifters. International Journal of Sports Medicine, 8 Suppl 1, 61-65.new window
Hameed, M., Orrell, R. W., Cobbold, M., Goldspink, G., & Harridge, S. D. (2003). Expression of IGF-I splice variants in young and old human skeletal muscle after high resistance exercise. Journal of Physiology, 547(Pt 1), 247-254.new window
Hansen, S., Kvorning, T., Kjaer, M., & Sjogaard, G. (2001). The effect of short-term strength training on human skeletal muscle: the importance of physiologically elevated hormone levels. Scandinavian Journal of Medicine and Science in Sports, 11(6), 347-354.
Henneman, E., Somjen, G., & Carpenter, D. O. (1965). Functional Significance of Cell Size in Spinal Motoneurons. Journal of Neurophysiology, 28, 560-580.
Hermansen, L., & Vaage, O. (1977). Lactate disappearance and glycogen synthesis in human muscle after maximal exercise. American Journal of Physiology, 233(5), E422-429.
Hoffman, J. R., Im, J., Rundell, K. W., Kang, J., Nioka, S., Spiering, B. A., et al. (2003). Effect of muscle oxygenation during resistance exercise on anabolic hormone response. Medicine and Science in Sports and Exercise, 35(11), 1929-1934.
Holm, L., van Hall, G., Rose, A. J., Miller, B. F., Doessing, S., Richter, E. A., et al. (2010). Contraction intensity and feeding affect collagen and myofibrillar protein synthesis rates differently in human skeletal muscle. American Journal of Physiology - Endocrinology and Metabolism, 298(2), E257-269.
Holz, M. K., Ballif, B. A., Gygi, S. P., & Blenis, J. (2005). mTOR and S6K1 mediate assembly of the translation preinitiation complex through dynamic protein interchange and ordered phosphorylation events. Cell, 123(4), 569-580.
Hornberger, T. A., & Esser, K. A. (2004). Mechanotransduction and the regulation of protein synthesis in skeletal muscle. Proceedings of the Nutrition Society, 63(2), 331-335.
Hornberger, T. A., Sukhija, K. B., & Chien, S. (2006). Regulation of mTOR by mechanically induced signaling events in skeletal muscle. Cell Cycle, 5(13), 1391-1396.
Houtman, C. J., Heerschap, A., Zwarts, M. J., & Stegeman, D. F. (2002). An additional phase in PCr use during sustained isometric exercise at 30% MVC in the tibialis anterior muscle. NMR in Biomedicine, 15(4), 270-277.
Houtman, C. J., Stegeman, D. F., Van Dijk, J. P., & Zwarts, M. J. (2003). Changes in muscle fiber conduction velocity indicate recruitment of distinct motor unit populations. Journal of Applied Physiology, 95(3), 1045-1054.
Hurley, B. F., & Roth, S. M. (2000). Strength training in the elderly: effects on risk factors for age-related diseases. Sports Medicine, 30(4), 249-268.
Jastrzebski, K., Hannan, K. M., Tchoubrieva, E. B., Hannan, R. D., & Pearson, R. B. (2007). Coordinate regulation of ribosome biogenesis and function by the ribosomal protein S6 kinase, a key mediator of mTOR function. Growth Factors, 25(4), 209-226.
Kadi, F. (2008). Cellular and molecular mechanisms responsible for the action of testosterone on human skeletal muscle. A basis for illegal performance enhancement. British Journal of Pharmacology, 154(3), 522-528.
Kahn, S. M., Hryb, D. J., Nakhla, A. M., Romas, N. A., & Rosner, W. (2002). Sex hormone-binding globulin is synthesized in target cells. Journal of Endocrinology, 175(1), 113-120.new window
Kawada, S., & Ishii, N. (2005). Skeletal muscle hypertrophy after chronic restriction of venous blood flow in rats. Medicine and Science in Sports and Exercise, 37(7), 1144-1150.
Kern, W., Dodt, C., Born, J., & Fehm, H. L. (1996). Changes in cortisol and growth hormone secretion during nocturnal sleep in the course of aging. The Journals of Gerontology. Series A, Biological Sciences and Medical Sciences, 51(1), M3-9.new window
Kimura, N., Tokunaga, C., Dalal, S., Richardson, C., Yoshino, K., Hara, K., et al. (2003). A possible linkage between AMP-activated protein kinase (AMPK) and mammalian target of rapamycin (mTOR) signalling pathway. Genes to Cells, 8(1), 65-79.new window
Klossner, S., Durieux, A. C., Freyssenet, D., & Flueck, M. (2009). Mechano transduction to muscle protein synthesis is modulated by FAK. European Journal of Applied Physiology and Occupational Physiology, 106(3), 389-398.
Koopman, R., Zorenc, A. H., Gransier, R. J., Cameron-Smith, D., & van Loon, L. J. (2006). Increase in S6K1 phosphorylation in human skeletal muscle following resistance exercise occurs mainly in type II muscle fibers. American Journal of Physiology. Endocrinology and Metabolism, 290(6), E1245-1252.
Kraemer, W. J. (1988). Endocrine responses to resistance exercise. Medicine and Science in Sports and Exercise, 20(5 Suppl), S152-157.
Kraemer, W. J., Adams, K., Cafarelli, E., Dudley, G. A., Dooly, C., Feigenbaum, M. S., et al. (2002). American College of Sports Medicine position stand. Progression models in resistance training for healthy adults. Medicine and Science in Sports and Exercise, 34(2), 364-380.
Kraemer, W. J., Aguilera, B. A., Terada, M., Newton, R. U., Lynch, J. M., Rosendaal, G., et al. (1995). Responses of IGF-I to endogenous increases in growth hormone after heavy-resistance exercise. Journal of Applied Physiology, 79(4), 1310-1315.
Kraemer, W. J., Fleck, S. J., & Evans, W. J. (1996). Strength and power training: physiological mechanisms of adaptation. Exercise and Sport Sciences Reviews, 24, 363-397.
Kraemer, W. J., Gordon, S. E., Fleck, S. J., Marchitelli, L. J., Mello, R., Dziados, J. E., et al. (1991). Endogenous anabolic hormonal and growth factor responses to heavy resistance exercise in males and females. International Journal of Sports Medicine, 12(2), 228-235.
Kraemer, W. J., Hakkinen, K., Newton, R. U., Nindl, B. C., Volek, J. S., McCormick, M., et al. (1999). Effects of heavy-resistance training on hormonal response patterns in younger vs. older men. Journal of Applied Physiology, 87(3), 982-992.
Kraemer, W. J., Marchitelli, L., Gordon, S. E., Harman, E., Dziados, J. E., Mello, R., et al. (1990). Hormonal and growth factor responses to heavy resistance exercise protocols. Journal of Applied Physiology, 69(4), 1442-1450.
Kraemer, W. J., & Ratamess, N. A. (2004). Fundamentals of resistance training: progression and exercise prescription. Medicine and Science in Sports and Exercise, 36(4), 674-688.
Kraemer, W. J., & Ratamess, N. A. (2005). Hormonal responses and adaptations to resistance exercise and training. Sports Medicine, 35(4), 339-361.
Kraemer, W. J., Staron, R. S., Hagerman, F. C., Hikida, R. S., Fry, A. C., Gordon, S. E., et al. (1998). The effects of short-term resistance training on endocrine function in men and women. European of Journal Applied Physiology Occupation Physiology, 78(1), 69-76.new window
Kraemer, W. J., Volek, J. S., Bush, J. A., Putukian, M., & Sebastianelli, W. J. (1998). Hormonal responses to consecutive days of heavy-resistance exercise with or without nutritional supplementation. Journal of Applied Physiology, 85(4), 1544-1555.
Kraemer, W. J., Volek, J. S., French, D. N., Rubin, M. R., Sharman, M. J., Gomez, A. L., et al. (2003). The effects of L-carnitine L-tartrate supplementation on hormonal responses to resistance exercise and recovery. Journal of Strength and Conditioning Research, 17(3), 455-462.
Kumar, V., Atherton, P., Smith, K., & Rennie, M. J. (2009). Human muscle protein synthesis and breakdown during and after exercise. Journal of Applied Physiology, 106(6), 2026-2039.
Kvorning, T., Andersen, M., Brixen, K., & Madsen, K. (2006). Suppression of endogenous testosterone production attenuates the response to strength training: a randomized, placebo-controlled, and blinded intervention study. American Journal of Physiology - Endocrinology and Metabolism, 291(6), E1325-1332.
Lamberts, S. W., van den Beld, A. W., & van der Lely, A. J. (1997). The endocrinology of aging. Science, 278(5337), 419-424.
Lassarre, C., Girard, F., Durand, J., & Raynaud, J. (1974). Kinetics of human growth hormone during submaximal exercise. Journal of Applied Physiology, 37(6), 826-830.
Le Roith, D., Bondy, C., Yakar, S., Liu, J. L., & Butler, A. (2001). The somatomedin hypothesis: 2001. Endocrine Reviews, 22(1), 53-74.new window
Leifke, E., Gorenoi, V., Wichers, C., Von Zur Muhlen, A., Von Buren, E., & Brabant, G. (2000). Age-related changes of serum sex hormones, insulin-like growth factor-1 and sex-hormone binding globulin levels in men: cross-sectional data from a healthy male cohort. Clinical Endocrinology, 53(6), 689-695.
Lim, C. T., Kola, B., & Korbonits, M. (2010). AMPK as a mediator of hormonal signalling. Journal of Molecular Endocrinology, 44(2), 87-97.
Lindinger, M. I., McKelvie, R. S., & Heigenhauser, G. J. (1995). K+ and Lac- distribution in humans during and after high-intensity exercise: role in muscle fatigue attenuation? Journal of Applied Physiology, 78(3), 765-777.
Loenneke, J. P., & Pujol, T. J. (2009). The Use of Occlusion Training to Produce Muscle Hypertrophy. Strength and Conditioning Journal, 31(3), 77-84
Loenneke, J. P., Wilson, G. J., & Wilson, J. M. (2010). A mechanistic approach to blood flow occlusion. International Journal of Sports Medicine, 31(1), 1-4.new window
Lu, S. S., Lau, C. P., Tung, Y. F., Huang, S. W., Chen, Y. H., Shih, H. C., et al. (1997). Lactate and the effects of exercise on testosterone secretion: evidence for the involvement of a cAMP-mediated mechanism. Medicine and Science in Sports and Exercise, 29(8), 1048-1054.
MacDougall, J. D., Gibala, M. J., Tarnopolsky, M. A., MacDonald, J. R., Interisano, S. A., & Yarasheski, K. E. (1995). The time course for elevated muscle protein synthesis following heavy resistance exercise. Canadian Journal of Applied Physiology, 20(4), 480-486.
Madarame, H., Neya, M., Ochi, E., Nakazato, K., Sato, Y., & Ishii, N. (2008). Cross-transfer effects of resistance training with blood flow restriction. Medicine and Science in Sports and Exercise, 40(2), 258-263.
Madarame, H., Sasaki, K., & Ishii, N. (2010). Endocrine responses to upper- and lower-limb resistance exercises with blood flow restriction. Acta Physiologica Hungarica, 97(2), 192-200.
Maresh, C. M., Cook, M. R., Cohen, H. D., Graham, C., & Gunn, W. S. (1988). Exercise testing in the evaluation of human responses to powerline frequency fields. Aviation Space and Environmental Medicine, 59(12), 1139-1145.
Matthiesson, K. L., & McLachlan, R. I. (2006). Male hormonal contraception: concept proven, product in sight? Human Reproduction Update, 12(4), 463-482.
McCall, G. E., Byrnes, W. C., Fleck, S. J., Dickinson, A., & Kraemer, W. J. (1999). Acute and chronic hormonal responses to resistance training designed to promote muscle hypertrophy. Canadina Journal of Applied Physiology, 24(1), 96-107.new window
McEwen, B. S. (1998). Protective and damaging effects of stress mediators. New England Journal of Medicine, 338(3), 171-179.
Miranda, L., Horman, S., De Potter, I., Hue, L., Jensen, J., & Rider, M. H. (2008). Effects of contraction and insulin on protein synthesis, AMP-activated protein kinase and phosphorylation state of translation factors in rat skeletal muscle. Pflugers Archiv. European Journal of Physiology, 455(6), 1129-1140.
Moritani, T., Sherman, W. M., Shibata, M., Matsumoto, T., & Shinohara, M. (1992). Oxygen availability and motor unit activity in humans. European Journal of Applied Physiology and Occupational Physiology, 64(6), 552-556.
Murthy, G., Hargens, A. R., Lehman, S., & Rempel, D. M. (2001). Ischemia causes muscle fatigue. Journal of Orthopaedic Research, 19(3), 436-440.
Nakagawa, E., Nagaya, N., Okumura, H., Enomoto, M., Oya, H., Ono, F., et al. (2002). Hyperglycaemia suppresses the secretion of ghrelin, a novel growth-hormone-releasing peptide: responses to the intravenous and oral administration of glucose. Clinical Science (London), 103(3), 325-328.
Nelson, M. E., Fiatarone, M. A., Morganti, C. M., Trice, I., Greenberg, R. A., & Evans, W. J. (1994). Effects of high-intensity strength training on multiple risk factors for osteoporotic fractures. A randomized controlled trial. Journal of the American Medical Association, 272(24), 1909-1914.
Nindl, B. C., Kraemer, W. J., Marx, J. O., Arciero, P. J., Dohi, K., Kellogg, M. D., et al. (2001). Overnight responses of the circulating IGF-I system after acute, heavy-resistance exercise. Journal of Applied Physiology, 90(4), 1319-1326.
Park, S., Kim, J. K., Choi, H. M., Kim, H. G., Beekley, M. D., & Nho, H. (2010). Increase in maximal oxygen uptake following 2-week walk training with blood flow occlusion in athletes. European Journal of Applied Physiology and Occupational Physiology, 109(4), 591-600.
Perry, S. R., Housh, T. J., Johnson, G. O., Ebersole, K. T., Bull, A. J., Evetovich, T. K., et al. (2001). Mechanomyography, electromyography, heart rate, and ratings of perceived exertion during incremental cycle ergometry. Journal of Sports Medicine and Physical Fitness, 41(2), 183-188.
Pescatello, L. S., Franklin, B. A., Fagard, R., Farquhar, W. B., Kelley, G. A., & Ray, C. A. (2004). American College of Sports Medicine position stand. Exercise and hypertension. Medicine and Science in Sports and Exercise, 36(3), 533-553.
Pierce, J. R., Clark, B. C., Ploutz-Snyder, L. L., & Kanaley, J. A. (2006). Growth hormone and muscle function responses to skeletal muscle ischemia. Journal of Applied Physiology, 101(6), 1588-1595.
Pollock, M. L., Franklin, B. A., Balady, G. J., Chaitman, B. L., Fleg, J. L., Fletcher, B., et al. (2000). AHA Science Advisory. Resistance exercise in individuals with and without cardiovascular disease: benefits, rationale, safety, and prescription: An advisory from the Committee on Exercise, Rehabilitation, and Prevention, Council on Clinical Cardiology, American Heart Association; Position paper endorsed by the American College of Sports Medicine. Circulation, 101(7), 828-833.
Pollock, M. L., Gaesser, G. A., Butcher, J. D., Després, J., Dishman, R. K., A., B., et al. (1998). American College of Sports Medicine Position Stand. The recommended quantity and quality of exercise for developing and maintaining cardiorespiratory and muscular fitness, and flexibility in healthy adults. Medicine and Science in Sports and Exercise, 30(6), 975-991.
Raastad, T., Bjoro, T., & Hallen, J. (2000). Hormonal responses to high- and moderate-intensity strength exercise. European Journal of Applied Physiology and Occupational Physiology, 82(1-2), 121-128.
Rasmussen, B. B. (2009). Phosphatidic acid: a novel mechanical mechanism for how resistance exercise activates mTORC1 signalling. Journal of Physiology, 587(Pt 14), 3415-3416.
Ratamess, N. A., Kraemer, W. J., Volek, J. S., Maresh, C. M., Vanheest, J. L., Sharman, M. J., et al. (2005). Androgen receptor content following heavy resistance exercise in men. Journal of Steroid Biochemistry and Molecular Biology, 93(1), 35-42.new window
Reeves, G. V., Kraemer, R. R., Hollander, D. B., Clavier, J., Thomas, C., Francois, M., et al. (2006). Comparison of hormone responses following light resistance exercise with partial vascular occlusion and moderately difficult resistance exercise without occlusion. Journal of Applied Physiology, 101(6), 1616-1622.
Rennie, M. J. (2003). Claims for the anabolic effects of growth hormone: a case of the emperor's new clothes? British Journal of Sports Medicine, 37(2), 100-105.
Rubin, M. R., Kraemer, W. J., Maresh, C. M., Volek, J. S., Ratamess, N. A., Vanheest, J. L., et al. (2005). High-affinity growth hormone binding protein and acute heavy resistance exercise. Medicine and Science in Sports and Exercise, 37(3), 395-403.
Sahlin, K., Harris, R. C., Nylind, B., & Hultman, E. (1976). Lactate content and pH in muscle obtained after dynamic exercise. Pflugers Archiv. European Journal of Physiology, 367(2), 143-149.
Schott, J., McCully, K., & Rutherford, O. M. (1995). The role of metabolites in strength training. II. Short versus long isometric contractions. European Journal of Applied Physiology and Occupational Physiology, 71(4), 337-341.
Sigal, R. J., Kenny, G. P., Wasserman, D. H., & Castaneda-Sceppa, C. (2004). Physical activity/exercise and type 2 diabetes. Diabetes Care, 27(10), 2518-2539.
Smilios, I., Pilianidis, T., Karamouzis, M., & Tokmakidis, S. P. (2003). Hormonal responses after various resistance exercise protocols. Medicine and Science in Sports and Exercise, 35(4), 644-654.
Spiering, B. A., Kraemer, W. J., Anderson, J. M., Armstrong, L. E., Nindl, B. C., Volek, J. S., et al. (2008). Resistance exercise biology: manipulation of resistance exercise programme variables determines the responses of cellular and molecular signalling pathways. Sports Medicine, 38(7), 527-540.
Spriet, L. L., Soderlund, K., Bergstrom, M., & Hultman, E. (1987). Skeletal muscle glycogenolysis, glycolysis, and pH during electrical stimulation in men. Journal of Applied Physiology, 62(2), 616-621.
Suga, T., Okita, K., Morita, N., Yokota, T., Hirabayashi, K., Horiuchi, M., et al. (2010). Dose effect on intramuscular metabolic stress during low-intensity resistance exercise with blood flow restriction. Journal of Applied Physiology, 108(6), 1563-1567.
Suga, T., Okita, K., Morita, N., Yokota, T., Hirabayashi, K., Horiuchi, M., et al. (2009). Intramuscular metabolism during low-intensity resistance exercise with blood flow restriction. Journal of Applied Physiology, 106(4), 1119-1124.
Takano, H., Morita, T., Iida, H., Asada, K., Kato, M., Uno, K., et al. (2005). Hemodynamic and hormonal responses to a short-term low-intensity resistance exercise with the reduction of muscle blood flow. European Journal of Applied Physiology and Occupational Physiology, 95(1), 65-73.new window
Takarada, Y., Nakamura, Y., Aruga, S., Onda, T., Miyazaki, S., & Ishii, N. (2000). Rapid increase in plasma growth hormone after low-intensity resistance exercise with vascular occlusion. Journal of Applied Physiology, 88(1), 61-65.new window
Takarada, Y., Sato, Y., & Ishii, N. (2002). Effects of resistance exercise combined with vascular occlusion on muscle function in athletes. European Journal of Applied Physiology and Occupational Physiology, 86(4), 308-314.
Takarada, Y., Takazawa, H., & Ishii, N. (2000). Applications of vascular occlusion diminish disuse atrophy of knee extensor muscles. Medicine and Science in Sports and Exercise, 32(12), 2035-2039.
Takarada, Y., Takazawa, H., Sato, Y., Takebayashi, S., Tanaka, Y., & Ishii, N. (2000). Effects of resistance exercise combined with moderate vascular occlusion on muscular function in humans. Journal of Applied Physiology, 88(6), 2097-2106.
Takarada, Y., Tsuruta, T., & Ishii, N. (2004). Cooperative effects of exercise and occlusive stimuli on muscular function in low-intensity resistance exercise with moderate vascular occlusion. Japanese Journal of Physiology, 54(6), 585-592.
Tannerstedt, J., Apro, W., & Blomstrand, E. (2009). Maximal lengthening contractions induce different signaling responses in the type I and type II fibers of human skeletal muscle. Journal of Applied Physiology, 106(4), 1412-1418.
Tarpenning, K. M., Wiswell, R. A., Hawkins, S. A., & Marcell, T. J. (2001). Influence of weight training exercise and modification of hormonal response on skeletal muscle growth. Journal of Science and Medicine in Sport, 4(4), 431-446.
Tremblay, M. S., Copeland, J. L., & Van Helder, W. (2004). Effect of training status and exercise mode on endogenous steroid hormones in men. Journal of Applied Physiology, 96(2), 531-539.
van Loon, L. J., Kruijshoop, M., Menheere, P. P., Wagenmakers, A. J., Saris, W. H., & Keizer, H. A. (2003). Amino acid ingestion strongly enhances insulin secretion in patients with long-term type 2 diabetes. Diabetes Care, 26(3), 625-630.
Vermeulen, A. (2000). Andropause. Maturitas, 34(1), 5-15.new window
Viru, M., Jansson, E., Viru, A., & Sundberg, C. J. (1998). Effect of restricted blood flow on exercise-induced hormone changes in healthy men. European Journal of Applied Physiology and Occupational Physiology, 77(6), 517-522.
Volek, J. S., Kraemer, W. J., Bush, J. A., Incledon, T., & Boetes, M. (1997). Testosterone and cortisol in relationship to dietary nutrients and resistance exercise. Journal of Applied Physiology, 82(1), 49-54.new window
Vollestad, N. K., & Blom, P. C. (1985). Effect of varying exercise intensity on glycogen depletion in human muscle fibres. Acta Physiologica Scandinavica, 125(3), 395-405.
Wang, X., & Proud, C. G. (2006). The mTOR pathway in the control of protein synthesis. Physiology (Bethesda), 21, 362-369.
Wang, P. S., Tsai, S. C., Hwang, G. S., Wang, S. W., Lu, C. C., Chen, J. J., et al. (1994). Calcitonin inhibits testosterone and luteinizing hormone secretion through a mechanism involving an increase in cAMP production in rats. Journal of Bone and Mineral Research, 9(10), 1583-1590.
Wernbom, M., Augustsson, J., & Raastad, T. (2008). Ischemic strength training: a low-load alternative to heavy resistance exercise? Scandinavian Journal of Medicine and Science in Sports, 18(4), 401-416.
West, D. W., Burd, N. A., Tang, J. E., Moore, D. R., Staples, A. W., Holwerda, A. M., et al. (2010). Elevations in ostensibly anabolic hormones with resistance exercise enhance neither training-induced muscle hypertrophy nor strength of the elbow flexors. Journal of Applied Physiology, 108(1), 60-67.new window

 
 
 
 
第一頁 上一頁 下一頁 最後一頁 top