Zeitschriftenartikel zum Thema „High intensity exercise“

Um die anderen Arten von Veröffentlichungen zu diesem Thema anzuzeigen, folgen Sie diesem Link: High intensity exercise.

Geben Sie eine Quelle nach APA, MLA, Chicago, Harvard und anderen Zitierweisen an

Wählen Sie eine Art der Quelle aus:

Machen Sie sich mit Top-50 Zeitschriftenartikel für die Forschung zum Thema "High intensity exercise" bekannt.

Neben jedem Werk im Literaturverzeichnis ist die Option "Zur Bibliographie hinzufügen" verfügbar. Nutzen Sie sie, wird Ihre bibliographische Angabe des gewählten Werkes nach der nötigen Zitierweise (APA, MLA, Harvard, Chicago, Vancouver usw.) automatisch gestaltet.

Sie können auch den vollen Text der wissenschaftlichen Publikation im PDF-Format herunterladen und eine Online-Annotation der Arbeit lesen, wenn die relevanten Parameter in den Metadaten verfügbar sind.

Sehen Sie die Zeitschriftenartikel für verschiedene Spezialgebieten durch und erstellen Sie Ihre Bibliographie auf korrekte Weise.

1

Maddigan, Meaghan E., Kathleen M. Sullivan, Israel Halperin, Fabien A. Basset und David G. Behm. „High tempo music prolongs high intensity exercise“. PeerJ 6 (08.01.2019): e6164. http://dx.doi.org/10.7717/peerj.6164.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Music has been shown to reduce rating of perceived exertion, increase exercise enjoyment and enhance exercise performance, mainly in low-moderate intensity exercises. However, the effects of music are less conclusive with high-intensity activities. The purpose of this with-participant design study was to compare the effects of high tempo music (130 bpm) to a no-music condition during repeated high intensity cycling bouts (80% of peak power output (PPO)) on the following measures: time to exercise end-point, rating of perceived exertion (RPE), heart rate (HR), breathing frequency, ventilatory kinetics and blood lactate (BL). Under the music condition, participants exercised 10.7% longer (p = 0.035; Effect size (ES) = 0.28) (increase of 1 min) and had higher HR (4%; p = 0.043; ES = 0.25), breathing frequency (11.6%; p < 0.001; ES = 0.57), and RER (7% at TTF; p = 0.021; ES = 1.1) during exercise, as measured at the exercise end-point. Trivial differences were observed between conditions in RPE and other ventilatory kinetics during exercise. Interestingly, 5 min post-exercise termination, HR recovery was 13.0% faster following the music condition (p < 0.05) despite that music was not played during this period. These results strengthen the notion that music can alter the association between central motor drive, central cardiovascular command and perceived exertion, and contribute to prolonged exercise durations at higher intensities along with a quicken HR recovery.
2

Mustafa Khan, Ghulam, Daniya Khan, Iram Saddiqa Aamir, Saher Ramees, Sassi Kanwal, Fatima Abid und Hina Moazzam. „HIGH INTENSITY EXERCISE;“. Professional Medical Journal 24, Nr. 03 (07.03.2017): 414–21. http://dx.doi.org/10.29309/tpmj/2017.24.03.1562.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Objectives: To clarify the interrelationship of high intensity exercise and durationof time on lipid profile, blood pressure and, body weight in young hyperlipidemic adults andto evaluate the beneficial effects on health in possible shorter duration of time. Study Design:Case control study. Setting: Department of Physiology, BMSI, JPMC, Karachi. Period: January2012 to April 2012. Methods: Total numbers of 60 volunteers moderately active, younghyperlipidemic adults, aged 30 to 50 years, thirty for each A (control), and B (high intensityexercise) group respectively were assigned to participate for 12 weeks exercise programwithout restricted caloric diet at pre–identified track of specific time. Lipids profile and age, sex,weight, BMI, blood pressure, and lab investigations like serum Total cholesterol, Triglycerides,LDL, and HDL, of each participant were recorded at baseline (0 day) and similarly on day30, 60 and 90 respectively. Results: At the end of the 12 weeks program in high intensityexercise group significant effects have been observed by lowering the level of total cholesterol,triglyceride (p <0.05) and lipoproteins variables, with increasing the level of HDL-c. While nosignificant change in blood pressure and weight reduction was observed between indicialand final measurement in control and high intensity exercise groups (p<o.o5). Conclusion:Finding of this study suggest that, high intensity exercise has positive effects on lipid profileby lowering the level of total cholesterol(TC), triglycerides(TGs) low density lipoprotein (LDL-c)and increasing the level of high density lipoprotein (HDL-c) in hyperlipidemic adults rapidlyin short duration of time, thus have beneficial effects by reducing the cardiac risk factors,chronic metabolic disorders and improving the cardio respiratory fitness but reduction in bloodpressure body weight can be achieved in short time by combination of high intensity exerciseand nutritional guidance of caloric restriction.
3

Pettitt, Robert W., und Ida E. Clark. „High-Intensity Exercise Tolerance“. Strength and Conditioning Journal 35, Nr. 2 (April 2013): 11–16. http://dx.doi.org/10.1519/ssc.0b013e31828a9520.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
4

Cooper, Simon B., Karah J. Dring und Mary E. Nevill. „High-Intensity Intermittent Exercise“. Current Sports Medicine Reports 15, Nr. 4 (2016): 245–51. http://dx.doi.org/10.1249/jsr.0000000000000273.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
5

Njapo, Steve Noutong, Brittney Heard und Mohamed Morsy. „HIGH INTENSITY EXERCISE INDUCED STEMI“. Journal of the American College of Cardiology 75, Nr. 11 (März 2020): 2446. http://dx.doi.org/10.1016/s0735-1097(20)33073-4.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
6

Roemmich, James, Kelsey Ufholz, Kyle Flack, Tori Kaster und William Siders. „High Intensity Interval Training to Increase Tolerance for Exercise Intensity“. Current Developments in Nutrition 4, Supplement_2 (29.05.2020): 1763. http://dx.doi.org/10.1093/cdn/nzaa066_018.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Abstract Objectives Increasing the reinforcing value of a stimulus occurs after repeated exposures to the reinforcer via neuroadaptations that increase the incentive salience of the stimulus. Exercise is a reinforcer and increasing exercise reinforcement (RRVex) may be dependent on simultaneously increasing tolerance for exercise intensity. Positive outcome expectancy (POE) of participating in an intervention can be an important determinant of treatment efficacy, such as when attempting to increase tolerance for exercise intensity or RRVex. We hypothesized that (1) high-intensity interval training (HIIT) that produces great discomfort would increase tolerance for exercise intensity, (2) adding a positive outcome expectancy (POE) component to HIIT would further increase tolerance for exercise intensity and, (3) increases in tolerance for exercise discomfort would mediate increases in RRVex. Methods A randomized controlled trial with a factorial design included HIIT + POE (n = 33 adults, n = 19 women) and HIIT-only (n = 33, n = 19 women) groups. Both groups participated in HIIT 3 d/wk for 6 wks. HIIT + POE received POE treatment each exercise session. Outcomes were measured at baseline, after 6 weeks of HIIT, and 4 weeks post-HIIT (10 wk). Changes in the RRVex were assessed by a progressive ratio schedule of reinforcement task. Other outcomes were outcome expectations, tolerance for exercise intensity, and behavior regulations of exercise. Results Outcome expectancy did not change in either group. Tolerance for exercise discomfort increased (P &lt; .001) above baseline by 12% at 6 wk and 13% at 10 wk. Intrinsic, integrated, and identified behavior regulations of exercise were all increased (P &lt; .01) at 6 wk and remained so at 10 wk. However, RRVex was not changed and change in RRVex was not correlated with change in tolerance for exercise intensity. Conclusions HIIT increases tolerance for exercise intensity and intrinsic, integrated, and identified behavior regulations of exercise. Funding Sources USDA-ARS.
7

Allen, David, und Håkan Westerblad. „What limits exercise during high-intensity aerobic exercise?“ European Journal of Applied Physiology 110, Nr. 3 (15.06.2010): 661–62. http://dx.doi.org/10.1007/s00421-010-1538-z.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
8

De Feo, P. „Is high-intensity exercise better than moderate-intensity exercise for weight loss?“ Nutrition, Metabolism and Cardiovascular Diseases 23, Nr. 11 (November 2013): 1037–42. http://dx.doi.org/10.1016/j.numecd.2013.06.002.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
9

Behm, David G., Dario Cappa und Geoffrey A. Power. „Trunk muscle activation during moderate- and high-intensity running“. Applied Physiology, Nutrition, and Metabolism 34, Nr. 6 (Dezember 2009): 1008–16. http://dx.doi.org/10.1139/h09-102.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Time constraints are cited as a barrier to regular exercise. If particular exercises can achieve multiple training functions, the number of exercises and the time needed to achieve a training goal may be decreased. It was the objective of this study to compare the extent of trunk muscle electromyographic (EMG) activity during running and callisthenic activities. EMG activity of the external obliques, lower abdominals (LA), upper lumbar erector spinae (ULES), and lumbosacral erector spinae (LSES) was monitored while triathletes and active nonrunners ran on a treadmill for 30 min at 60% and 80% of their maximum heart rate (HR) reserve, as well as during 30 repetitions of a partial curl-up and 3 min of a modified Biering-Sørensen back extension exercise. The mean root mean square (RMS) amplitude of the EMG signal was monitored over 10-s periods with measures normalized to a maximum voluntary contraction rotating curl-up (external obliques), hollowing exercise (LA), or back extension (ULES and LSES). A main effect for group was that triathletes had greater overall activation of the external obliques (p < 0.05), LA (p = 0.01), and LSES (p < 0.05) than did nonrunners. Main effects for exercise type showed that the external obliques had less EMG activity during 60% and 80% runs, respectively, than with the curl-ups (p = 0.001). The back extension exercise provided less ULES (p = 0.009) and LSES (p = 0.0001) EMG activity than the 60% and 80% runs, respectively. In conclusion, triathletes had greater trunk activation than nonrunners did while running, which could have contributed to their better performance. Back-stabilizing muscles can be activated more effectively with running than with a prolonged back extension activity. Running can be considered as an efficient, multifunctional exercise combining cardiovascular and trunk endurance benefits.
10

Goršič, Maja, Imre Cikajlo, Nika Goljar und Domen Novak. „A Multisession Evaluation of a Collaborative Virtual Environment for Arm Rehabilitation“. PRESENCE: Virtual and Augmented Reality 27, Nr. 3 (Juli 2020): 274–86. http://dx.doi.org/10.1162/pres_a_00331.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
In recent years, several multi-user virtual environments (VEs) have been developed to promote motivation and exercise intensity in motor rehabilitation. While competitive VEs have been extensively evaluated, collaborative and competitive rehabilitation VEs have seen relatively little study. Therefore, this article presents an evaluation of a VE for post-stroke arm rehabilitation that mimics everyday kitchen tasks and can be used either solo or collaboratively. Twenty subacute stroke survivors exercised with the VE for four sessions, with the first and third sessions involving solo exercise and the other two involving collaborative exercise. Exercise intensity was measured using inertial sensors while motivation was measured with questionnaires. Results showed high motivation and exercise intensity over all four sessions, and 11 of 20 participants preferred collaborative over solo exercise while only 4 preferred solo exercise. However, there were no differences in motivation, exercise duration, or exercise intensity between solo and collaborative sessions. Thus, we cannot currently claim that collaborative exercises are beneficial for upper limb rehabilitation. Future studies should evaluate other collaborative VE designs in different settings (e.g., at home) and with different participant pairs (e.g., patient-unimpaired) to find effective ways to utilize collaborative exercises in motor rehabilitation.
11

Gordon, Brett A., Caroline J. Taylor, Jarrod E. Church und Stephen D. Cousins. „A Comparison of the Gluco-Regulatory Responses to High-Intensity Interval Exercise and Resistance Exercise“. International Journal of Environmental Research and Public Health 18, Nr. 1 (02.01.2021): 287. http://dx.doi.org/10.3390/ijerph18010287.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
High-intensity interval exercise and resistance exercise both effectively lower blood glucose; however, it is not clear whether different regulatory mechanisms exist. This randomised cross-over study compared the acute gluco-regulatory and the physiological responses of high-intensity interval exercise and resistance exercise. Sixteen (eight males and eight females) recreationally active individuals, aged (mean ± SD) 22 ± 7 years, participated with a seven-day period between interventions. The high-intensity interval exercise trial consisted of twelve, 30 s cycling intervals at 80% of peak power capacity and 90 s active recovery. The resistance exercise trial consisted of four sets of 10 repetitions for three lower-limb exercises at 80% 1-RM, matched for duration of high-intensity interval exercise. Exercise was performed after an overnight fast, with blood samples collected every 30 min, for two hours after exercise. There was a significant interaction between time and intervention for glucose (p = 0.02), which was, on average (mean ± SD), 0.7 ± 0.7 mmol∙L−1 higher following high-intensity interval exercise, as compared to resistance exercise. Cortisol concentration over time was affected by intervention (p = 0.03), with cortisol 70 ± 103 ng∙mL−1 higher (p = 0.015), on average, following high-intensity interval exercise. Resistance exercise did not induce the acute rise in glucose that was induced by high-intensity interval exercise and appears to be an appropriate alternative to positively regulate blood glucose.
12

Lambert, Charles P., und Michael G. Flynn. „Fatigue during High-Intensity Intermittent Exercise“. Sports Medicine 32, Nr. 8 (2002): 511–22. http://dx.doi.org/10.2165/00007256-200232080-00003.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
13

Kastner, S., C. Becker und U. Lindemann. „High Intensity Functional Exercise (HIFE) Training“. physioscience 13, Nr. 03 (September 2017): 109–16. http://dx.doi.org/10.1055/s-0035-1567214.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Zusammenfassung Hintergrund: Das in Schweden entwickelte und evaluierte Trainingsprogramm „High Intensity Functional Exercise“ (HIFE) wird dort zur Sturzprävention im stationären Seniorenpflegebereich eingesetzt. Das Programm besteht aus funktionellen, hochintensiven Kraft- und Balanceübungen und könnte eine wertvolle Ergänzung zu bereits bestehenden Sturzpräventionsprogrammen in Deutschland darstellen. Ziel: Überprüfung der Anwendbarkeit des HIFE-Programms in 2 deutschen Pflegeheimen. Methode: Machbarkeitsstudie einer 10-wöchigen HIFE-Intervention mit 16 Teilnehmern*. Primäre Zielparameter waren Teilnahme- und Drop-out-Raten, Beschwerden und Nebenwirkungen, Akzeptanz bei Heimleitung und Pflegepersonal. Sekundäre Zielparameter umfassten 30-Seconds Chair Rising Test (30 CR), 5-Seconds Chair Rising Test (5 CR), habituelle und maximale Gehgeschwindigkeit, Timed Up and Go Test (TUG), Sit and Reach Test (SR) und Standbalance zur Beschreibung funktioneller Veränderungen im Prä-Post-Vergleich. Ergebnisse: Während der Intervention traten keine schwerwiegenden Nebenwirkungen auf. Teilnahmeraten und Akzeptanz des HIFE-Programms waren bei Heimleitung und Pflegepersonal sehr gut und die Drop-out-Rate gering. Die Standbalance (p < 0,05), funktionelle Beinkraft (30 CR: p < 0,005, 5 CR: p < 0,005) und habituelle Gehgeschwindigkeit (p < 0,05) verbesserten sich signifikant. Die Verbesserungen der maximalen Gehgeschwindigkeit (p = 0,075) und im TUG (p = 0,05) erreichten kein signifikantes Niveau. Die Beweglichkeit beim SR veränderte sich nicht (p = 0,608). Schlussfolgerungen: Das HIFE-Trainingsprogramm lässt sich ist in Deutschland sicher und praktikabel anwenden. Verbesserungen funktioneller Parameter sind zu erwarten, was insgesamt einen wertvollen Beitrag zur Sturzprävention in Seniorenheimen leisten kann.
14

Mitchell, Joel B., Melissa M. Rogers, John T. Basset und Kimberly A. Hubing. „Fatigue During High-Intensity Endurance Exercise“. Journal of Strength and Conditioning Research 28, Nr. 7 (Juli 2014): 1906–14. http://dx.doi.org/10.1519/jsc.0000000000000319.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
15

Coyle, Edward F., Joel D. Trinity, Matthew D. Pahnke, Joshua F. Lee und Kenneth C. Beck. „Cardiac Output During High Intensity Exercise“. Medicine & Science in Sports & Exercise 42 (Mai 2010): 4. http://dx.doi.org/10.1249/01.mss.0000384238.27287.b8.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
16

KOMIYAMA, TAKAAKI, YUKIYA TANOUE, MIZUKI SUDO, JOSEPH T. COSTELLO, YOSHINARI UEHARA, YASUKI HIGAKI und SOICHI ANDO. „Cognitive Impairment during High-Intensity Exercise“. Medicine & Science in Sports & Exercise 52, Nr. 3 (März 2020): 561–68. http://dx.doi.org/10.1249/mss.0000000000002183.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
17

Lagally, Kristen M., Kara I. Gallagher, Robert J. Robertson, Randall Gearhart und Fredric L. Goss. „Ratings of Perceived Exertion during Low- and High-Intensity Resistance Exercise by Young Adults“. Perceptual and Motor Skills 94, Nr. 3 (Juni 2002): 723–31. http://dx.doi.org/10.2466/pms.2002.94.3.723.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Ratings of perceived exertion (RPE) are commonly used to monitor the intensity of aerobic exercise. Whether ratings of perceived exertion can be used similarly during resistance exercise is unclear. To examine this question, perceived exertion was measured at 30% and 90% of the one-repetition maximum (1-RM), while holding work constant between intensities. Ratings for the active muscles and for the overall body were examined during both intensities. 10 male (age = 23.2 ± 3.6 yr.) and nine female (age = 21.8 ± 2.7 yr.) volunteers underwent a one-repetition maximum procedure for each of the following exercises; bench press, leg press, latissimus pull down, triceps press, biceps curl, shoulder press, and calf raise. All subjects then completed two experimental trials on separate days. The high-intensity trial consisted of one set of five repetitions at 90% of the one-repetition maximum. The low-intensity trial consisted of one set of 15 repetitions at 30% of the one-repetition maximum. Active muscle and overall body ratings of perceived exertion were obtained immediately at termination of each of the seven exercises at both intensities. A two-factor (RPE x Intensity) repeated-measures analysis of variance was performed separately for each exercise. Both active muscle and overall body ratings of perceived exertion were higher ( p<.01) for the high-intensity trial than for the low intensity trial. Active muscle ratings were higher ( p<.01) than overall body ratings for all exercises. Ratings of perceived exertion during resistance exercise are related to intensity of the resistance exercise (percentage of the one-repetition maximum). This information suggests that ratings of perceived exertion can provide information regarding the intensity of resistance exercise. Furthermore, sensations of exertion in the active muscles during resistance exercise are greater than sensations for the overall body.
18

Mahmoud, Tarek H., Mona Sayed Ahmed, Haitham M. Saleh, Nadia Mohamed Abdelhakiem, Raafat Ezzeldeen Abd Elhameed und Mohamed Ahmed Gad Allah. „Effect of High-intensity Aerobic Exercise Versus Moderate-intensity Aerobic Exercise on Serum Uric Acid in Hypertension“. NeuroQuantology 20, Nr. 4 (30.04.2022): 552–63. http://dx.doi.org/10.14704/nq.2022.20.4.nq22336.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Background & purpose: Methods: Thirty hypertensive men (Mild Essential Hypertension) with ages ranging from 45 to 55 years old and BMIs ranging from 25 to 35 kg/m2 took part in the study (overweight and class1 obesity). The participants were classified randomly into two main domains; A and B. As for the first Group, (A) did high-intensity (interval) aerobic activity, while group (B) did moderate-intensity (continuous) aerobic exercise. Both categories exercised for 30-40 minutes on an electrical treadmill three times a week for eight weeks. Blood uric acid, serum creatinine, and serum urea were tested for each patient before and after the trial. The arterial blood pressure measurements were also taken before and after the study to draw comparisons and to set evaluations. Results: Both high-intensity interval and continuous moderate exercise had a significant lowering effect on serum uric acid and creatinine (p.05), with continuous moderate exercise having a better effect with a 16 percent improvement in serum uric acid and a 27.5 percent improvement in creatinine, while high-intensity interval exercise had an 8 percent improvement in serum uric acid and a 17 percent improvement in creatinine. On the other hand, arterial blood pressure and blood urea levels improved at the same rate. Conclusion: Moderate continuous exercise reduces serum uric acid and creatinine better than high-intensity interval exercise. Both have the same effect on blood pressure and urea, but moderate continuous exercise has a crystal-clear effect that cannot be denied.
19

Safarinejad, Mohammad Reza, Kamran Azma und Ali Asgar Kolahi. „The effects of intensive, long-term treadmill running on reproductive hormones, hypothalamus–pituitary–testis axis, and semen quality: a randomized controlled study“. Journal of Endocrinology 200, Nr. 3 (03.12.2008): 259–71. http://dx.doi.org/10.1677/joe-08-0477.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Effects of intensive exercise on hypothalamus–pituitary–testis (HPT) axis remain controversial. Our aim was to determine the effects of intensive, long-term treadmill running on reproductive hormones, HPT axis, and semen quality. A total of 286 subjects were randomly assigned to moderate-intensity exercise (∼60% maximal oxygen uptake (VO2max); group 1, n=143) and high-intensity exercise (∼80% VO2max; group 2, n=143) groups. The two groups exercised for 60 weeks in five sessions per week, each session lasting 120 min. This was followed by a 36-week low-intensity exercise recovery period. All subjects underwent routine semen analysis. Blood samples were drawn for the determination of the levels of the following hormones: LH, FSH, prolactin, testosterone (T), free testosterone (fT), inhibin B, and sex hormone-binding globulin (SHBG). The HPT axis was assessed using GnRH and human chorionic gonadotropin tests. After 24 weeks of exercise, the subjects exercising with high intensity demonstrated significantly declined semen parameters compared with those exercising with moderate intensity (P=0.03). Serum T and fT began to decrease, and serum SHBG began to increase at the end of 12 weeks with both moderate- and high-intensity exercises. The serum LH and FSH concentrations decreased below the baseline level at 12 weeks in both groups (P=0.07 in group 1 and 0.03 in group 2). Both groups had blunted LH and FSH responses to GnRH. These parameters improved to their pre-exercise level during the recovery period. Long-term strenuous treadmill exercises (overtraining syndrome) have a deleterious effect on reproduction.
20

Hwang, Ji Sun, Tae Young Kim, Moon-Hyon Hwang und Won Jun Lee. „Exercise and Neuroplasticity: Benefits of High Intensity Interval Exercise“. Journal of Life Science 26, Nr. 1 (30.01.2016): 129–39. http://dx.doi.org/10.5352/jls.2016.26.1.129.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
21

Navalta, James W., und Stephen P. Hrncir. „Core Stabilization Exercises Enhance Lactate Clearance Following High-Intensity Exercise“. Journal of Strength and Conditioning Research 21, Nr. 4 (2007): 1305. http://dx.doi.org/10.1519/r-21546.1.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
22

NAVALTA, JAMES W., und STEPHEN P. HRNCIR. „CORE STABILIZATION EXERCISES ENHANCE LACTATE CLEARANCE FOLLOWING HIGH-INTENSITY EXERCISE“. Journal of Strength and Conditioning Research 21, Nr. 4 (November 2007): 1305–9. http://dx.doi.org/10.1519/00124278-200711000-00057.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
23

Andrews, Sophie C., Dylan Curtin, Ziarih Hawi, Jaeger Wongtrakun, Julie C. Stout und James P. Coxon. „Intensity Matters: High-intensity Interval Exercise Enhances Motor Cortex Plasticity More Than Moderate Exercise“. Cerebral Cortex 30, Nr. 1 (01.05.2019): 101–12. http://dx.doi.org/10.1093/cercor/bhz075.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Abstract A single bout of cardiovascular exercise can enhance plasticity in human cortex; however, the intensity required for optimal enhancement is debated. We investigated the effect of exercise intensity on motor cortex synaptic plasticity, using transcranial magnetic stimulation. Twenty healthy adults (Mage = 35.10 ± 13.25 years) completed three sessions. Measures of cortico-motor excitability (CME) and inhibition were obtained before and after a 20-min bout of either high-intensity interval exercise, moderate-intensity continuous exercise, or rest, and again after intermittent theta burst stimulation (iTBS). Results showed that high-intensity interval exercise enhanced iTBS plasticity more than rest, evidenced by increased CME and intracortical facilitation, and reduced intracortical inhibition. In comparison, the effect of moderate-intensity exercise was intermediate between high-intensity exercise and rest. Importantly, analysis of each participant’s plasticity response profile indicated that high-intensity exercise increased the likelihood of a facilitatory response to iTBS. We also established that the brain-derived neurotrophic factor Val66Met polymorphism attenuated plasticity responses following high-intensity exercise. These findings suggest that high-intensity interval exercise should be considered not only when planning exercise interventions designed to enhance neuroplasticity, but also to maximize the therapeutic potential of non-invasive brain stimulation. Additionally, genetic profiling may enhance efficacy of exercise interventions for brain health.
24

Hargreaves, Mark, Michael J. McKenna, David G. Jenkins, Stuart A. Warmington, Jia L. Li, Rodney J. Snow und Mark A. Febbraio. „Muscle metabolites and performance during high-intensity, intermittent exercise“. Journal of Applied Physiology 84, Nr. 5 (01.05.1998): 1687–91. http://dx.doi.org/10.1152/jappl.1998.84.5.1687.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Six men were studied during four 30-s “all-out” exercise bouts on an air-braked cycle ergometer. The first three exercise bouts were separated by 4 min of passive recovery; after the third bout, subjects rested for 4 min, exercised for 30 min at 30–35% peak O2 consumption, and rested for a further 60 min before completing the fourth exercise bout. Peak power and total work were reduced ( P < 0.05) during bout 3 [765 ± 60 (SE) W; 15.8 ± 1.0 kJ] compared with bout 1 (1,168 ± 55 W, 23.8 ± 1.2 kJ), but no difference in exercise performance was observed between bouts 1 and 4 (1,094 ± 64 W, 23.2 ± 1.4 kJ). Before bout 3, muscle ATP, creatine phosphate (CP), glycogen, pH, and sarcoplasmic reticulum (SR) Ca2+ uptake were reduced, while muscle lactate and inosine 5′-monophosphate were increased. Muscle ATP and glycogen before bout 4 remained lower than values before bout 1( P < 0.05), but there were no differences in muscle inosine 5′-monophosphate, lactate, pH, and SR Ca2+ uptake. Muscle CP levels before bout 4 had increased above resting levels. Consistent with the decline in muscle ATP were increases in hypoxanthine and inosine before bouts 3 and 4. The decline in exercise performance does not appear to be related to a reduction in muscle glycogen. Instead, it may be caused by reduced CP availability, increased H+ concentration, impairment in SR function, or some other fatigue-inducing agent.
25

Ko, Young-woo, Sun Mi Kim, Kyoung Doo Kang und Doug Hyun Han. „Changes in Functional Connectivity Between Default Mode Network and Attention Network in Response to Changes in Aerobic Exercise Intensity“. Psychiatry Investigation 20, Nr. 1 (25.01.2023): 27–34. http://dx.doi.org/10.30773/pi.2022.0245.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Objective Aerobic exercise may be associated with changes in brain activity within the default mode network (DMN) and dorsal attention network (DAN). We hypothesized that changes in functional connectivity (FC) within the DMN and DAN might be most effectively activated by moderate-intensity exercise.Methods Resting-state functional magnetic resonance imaging scans and visuospatial attention tests after resting were performed before and after each of moderate- and high-intensity aerobic exercises (10 min each) in 15 healthy male volunteers.Results The reaction time during the attention test increased significantly, and the rate of correct responses decreased from moderate-intensity exercise condition to high-intensity exercise condition. FC within the DMN under high-intensity exercise condition was higher than that under pre-exercise and moderate-intensity exercise conditions. FC within the DAN under moderate-intensity exercise condition was the highest, whereas FC between the DMN and DAN under moderate-intensity exercise condition was the lowest. Changes in cognitive domain functions were associated with changes in FC between the DMN and DAN.Conclusion Our results support the inverted-U hypothesis of maximum arousal efficacy during moderate exercise. Both cognitive domains, namely, the attention system and brain activity domains, may be better under moderate-intensity exercise than under high-intensity exercise.
26

Badenhop, Dalynn T., Meghan M. Long, C. Matt Laurent und K. Todd Keylock. „High-Intensity Interval Versus Moderate-Intensity Continuous Training in Cardiac Rehabilitation“. Journal of Clinical Exercise Physiology 9, Nr. 1 (01.03.2020): 10–16. http://dx.doi.org/10.31189/2165-6193-9.1.10.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
ABSTRACT Background: Past research has compared the effects of moderate-intensity continuous training (MCT) versus high-intensity interval training (HIIT) in phase 2 cardiac rehabilitation patients, but with conflicting results. Therefore, the purpose of this study was to evaluate if HIIT leads to greater improvements in functional capacity when compared with MCT in a group of phase 2 cardiac rehabilitation patients. Methods: Eighteen patients in a phase 2 cardiac rehabilitation program completed precardiopulmonary and postcardiopulmonary exercise tests, a 12-min walk test (12MWT), and resting blood pressure (BP). After 2 weeks of run-in, patients were randomly assigned to 10 weeks of HIIT (alternating periods of 80%–90% heart rate [HR] reserve and 60%–70% HR reserve) or MCT (60%–80% HR reserve) exercise group. Changes in VO2 peak, 12MWT distance, and BP (mm Hg) were analyzed by independent t test. Results: The average patient was 65 years old, 1.75 m tall, and overweight. VO2 peak values improved for individuals in both exercise modalities. There was no significant difference between the exercise groups (P = 0.174). In addition, both groups improved their 12MWT distance, resting systolic, and diastolic BP (DBP), with no significant difference in improvements between the 2 exercise groups. Conclusion: In this study, HIIT was not more effective than MCT for improving functional capacity in a group of phase 2 cardiac rehabilitation patients. However, since HIIT was equally effective compared with MCT in several measures, it provides another option for exercise prescription to the traditional prescription for this population.
27

Gearhart, Randall F., M. Daniel Becque, Chad M. Palm und Matthew D. Hutchins. „Rating Perceived Exertion during Short Duration, Very High Intensity Cycle Exercise“. Perceptual and Motor Skills 100, Nr. 3 (Juni 2005): 767–73. http://dx.doi.org/10.2466/pms.100.3.767-773.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
This study compared undifferentiated ratings of perceived exertion (RPE) during short duration, very high intensity cycle exercise using high and low resistance. 30 recreationally trained males (24.2 ± 2.4 yr.) were memory-anchored to the Borg 15-category scale. The high and low resistance exercises were defined by 30-sec. maximum tests assigned in counterbalanced order, with resistances set before testing during an orientation session. High resistance was 10% of body mass. Low resistance resulted in the same total work as the high resistance over the 30-sec. sessions (± 5%) but increased pedal rate. RPE was taken at 8, 13, 18, 23, and 28 sec. during the high and the low resistance exercises. Measurements were compared using a 2-way repeated-measures analysis of variance. RPE was significantly greater ( p = .005) for the high than the low resistance exercise at each interval. RPE increased when the subjects were required to pedal against a greater resistance and produce the highest forces. These RPE data are consistent with data from both aerobic cycle and resistance exercise. The data suggest that instantaneous force production, not summed work, is a primary determinant of RPE. All of these observations support Cafarelli's theoretical model of effort sense. In conclusion, as an individual generates more force during high resistance exercise than in light resistance exercise, a potential explanation of our results is that the increased motor outflow and corollary sensory signal lead to a greater sense of effort.
28

Rodrigues, Alesson, und Leonardo De Lucca. „Acute leptin response after high intensity interval and moderate intensity continuous runs“. European Journal of Human Movement 45 (2020): 26–35. http://dx.doi.org/10.21134/eurjhm.2020.45.3.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
The possible direct role of exercise intensity and duration on leptin concentrations is conflicting. The aim of this study was to evaluate the acute effects of high intensity interval (HIIE) and moderate intensity continuous (MICE) exercise on plasma leptin response. Seven young volunteers underwent three tests: 1) a treadmill graded exercise test to identify running peak velocity (PV); 2) HIIE: 5 × 2 min work bouts at 90% of PV, interspersed by 2 min of passive recovery and; 3) MICE: 30 min at 70 % of PV. Blood samples were drawn for the assays of leptin before and 30 minutes after HIIE and MICE. A 2-way repeated measures ANOVA showed a significant main effect of time [F(1,6) =17,52; p=0,006], no significant effect of condition (type of exercise) (F(1,6) = 0,16; p = 0,68) and no significant interaction (condition × time) (F(1,6)= 0,48, p=0,51). Leptin decreased 30 min after HIIE (t= 2,95, p=0,025) and MICE (t=4,18; p=0,005). There was no difference between the HIIE and MICE conditions immediately after exercise (t=0,90; p=0,40). After HIIE and MICE, leptin decreased in the same magnitude. It appears that both exercise modalities result in physical stress which is sufficient to improve short-term leptin sensibility.
29

Wang, Chun-Chih, Chien-Heng Chu, I.-Hua Chu, Kuei-Hui Chan und Yu-Kai Chang. „Executive Function During Acute Exercise: The Role of Exercise Intensity“. Journal of Sport and Exercise Psychology 35, Nr. 4 (August 2013): 358–67. http://dx.doi.org/10.1123/jsep.35.4.358.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
This study was designed to examine the modulation of executive functions during acute exercise and to determine whether exercise intensity moderates this relationship. Eighty college-aged adults were recruited and randomly assigned into one of the four following groups: control, 30%, 50%, and 80% heart rate reserve. The Wisconsin Card Sorting Test (WCST) was administered during each intervention. The results indicated that the majority of the WCST performances were impaired in the high exercise intensity group relative to those of the other three groups, whereas similar performance rates were maintained in the low- and moderate-intensity groups. These findings suggest that transient hypofrontality occurs during high-intensity exercise, but not during low- and moderate-intensity exercises. Future research aimed at employing the dual-mode theory, and applying the reticular-activating hypofrontality model is recommended to further the current knowledge.
30

Goto, K., S. Nagasawa, O. Yanagisawa, F. Kaneko, T. Kizuka und K. Takamatsu. „ADDITION OF LOW INTENSITY RESISTANCE EXERCISE TO HIGH INTENSITY RESISTANCE EXERCISE INCREASES MUSCULAR STRENGTH.“ Medicine & Science in Sports & Exercise 34, Nr. 5 (Mai 2002): S199. http://dx.doi.org/10.1097/00005768-200205001-01122.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
31

Thompson, Jenna, Matthew Wolfe, Meral Culver, Kelly E. Johnson und Justin P. Guilkey. „Exercise Intensity, Energy Expenditure And Enjoyment During Variable High Intensity Exercise In Healthy Adults.“ Medicine & Science in Sports & Exercise 51, Supplement (Juni 2019): 327. http://dx.doi.org/10.1249/01.mss.0000561489.35384.2a.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
32

Pimenta, Flávia C., Fábio Tanil Montrezol, Victor Zuniga Dourado, Luís Fernando Marcelino da Silva, Gabriela Alves Borba, Wesley de Oliveira Vieira und Alessandra Medeiros. „High-intensity interval exercise promotes post-exercise hypotension of greater magnitude compared to moderate-intensity continuous exercise“. European Journal of Applied Physiology 119, Nr. 5 (08.03.2019): 1235–43. http://dx.doi.org/10.1007/s00421-019-04114-9.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
33

Järvinen, Laura, Sofi Lundin Petersdotter und Thomas Chaillou. „High-intensity resistance exercise is not as effective as traditional high-intensity interval exercise for increasing the cardiorespiratory response and energy expenditure in recreationally active subjects“. European Journal of Applied Physiology 122, Nr. 2 (19.11.2021): 459–74. http://dx.doi.org/10.1007/s00421-021-04849-4.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Abstract Purpose Traditional high-intensity interval exercise (HIIE) highly stimulates the cardiorespiratory system and increases energy expenditure (EE) during exercise. High-intensity resistance exercise (HIRE) has become more popular in recreationally active subjects. The physiological responses to HIRE performed with light or moderate load is currently largely unknown. Here, we examined the effect of the type of interval exercise [HIRE at 40% (HIRE40) and 60% (HIRE60) 1-RM vs. traditional HIIE] on the cardiorespiratory response and EE during and after exercise. Methods Fifteen recreationally active adults randomly completed traditional HIIE on an ergocyle, HIRE40 and HIRE60. The sessions consisted of two sets of ten 30-s intervals (power at 100% VO2max during HIIE; maximal number of repetitions for 10 different free-weight exercises during HIRE40 and HIRE60) separated by 30-s active recovery periods. Gas exchange, heart rate (HR) and EE were assessed during and after exercise. Results VO2mean, VO2peak, HRmean, the time spent above 90% VO2max and HRmax, and aerobic EE were lower in both HIRE sessions compared with HIIE (P < 0.05). Anaerobic glycolytic contribution to total exercise EE was higher in HIRE40 and HIRE60 compared with HIIE (P < 0.001). EE from excess post-exercise oxygen consumption (EPOC) was similar after the three sessions. Overall, similar cardiorespiratory responses and EE were found in HIRE40 and HIRE60. Conclusions HIRE is not as effective as HIIE for increasing the cardiorespiratory response and EE during exercise, while EPOC remains similar in HIRE and HIIE. These parameters are not substantially different between HIRE40 and HIRE60.
34

Moriarty, Terence, Kelsey Bourbeau, Bryanne Bellovary und Micah N. Zuhl. „Exercise Intensity Influences Prefrontal Cortex Oxygenation during Cognitive Testing“. Behavioral Sciences 9, Nr. 8 (26.07.2019): 83. http://dx.doi.org/10.3390/bs9080083.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
Activation changes in the prefrontal cortex (PFC) regions have been linked to acute exercise-induced improvements in cognitive performance. The type of exercise performed may influence PFC activation, and further impact cognitive function. The present study aimed to compare PFC activation during cognitive testing after moderate-intensity, high intensity, and yoga exercises, and to determine if PFC activation is linked to cognitive performance. Eight subjects (four male and four female), aged 35 ± 5 completed a control, high intensity, moderate intensity, and yoga exercises followed by administration of a cognitive task (NIH Toolbox Fluid Cognition). Left and right PFC activation (LPFC and RPFC, respectively) were evaluated by measuring hemoglobin difference (Hbdiff) changes during post-exercise cognitive assessment using functional near infrared spectroscopy (fNIRS). Activation during the cognitive test was higher in the LPFC after moderate intensity exercise compared to control, high intensity, and yoga (5.30 ± 6.65 vs. 2.26 ± 2.40, 2.50 ± 1.48, 2.41 ± 2.36 μM, p < 0.05, respectively). A negative relationship was detected between LPFC and processing speed after exercise. PFC activation did not align with cognitive performance. However, acute exercise, regardless of type, appeared to alter neural processing. Specifically, less PFC activation was required for a given neural output after exercise.
35

Vargas, Nicole T., Christopher L. Chapman, Blair D. Johnson, Rob Gathercole und Zachary J. Schlader. „Exercise intensity independently modulates thermal behavior during exercise recovery but not during exercise“. Journal of Applied Physiology 126, Nr. 4 (01.04.2019): 1150–59. http://dx.doi.org/10.1152/japplphysiol.00992.2018.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
We tested the hypothesis that thermal behavior is greater during and after high- compared with moderate-intensity exercise. In a 27°C, 20% relative humidity environment, 20 participants (10 women, 10 men) cycled for 30 min at moderate [53% (SD 6) peak oxygen uptake (V̇o2peak) or high [78% (SD 6) V̇o2peak] intensity, followed by 120 min of recovery. Mean skin and core temperatures and mean skin wettedness were recorded continuously. Participants maintained thermally comfortable neck temperatures with a custom-made neck device. Neck device temperature provided an index of thermal behavior. The weighted average of mean skin and core temperatures and mean skin wettedness provided an indication of the afferent stimulus to thermally behave. Mean skin and core temperatures were greater at end-exercise in high intensity ( P < 0.01). Core temperature remained elevated in high intensity until 70 min of recovery ( P = 0.03). Mean skin wettedness and the afferent stimulus were greater at 10–20 min of exercise in high intensity ( P ≤ 0.03) and remained elevated until 60 min of recovery ( P < 0.01). Neck device temperature was lower during exercise in high versus moderate intensity ( P ≤ 0.02). There was a strong relation between the afferent stimulus and neck device temperature during exercise (high: R2 = 0.82, P < 0.01; moderate: R2 = 0.95, P < 0.01) and recovery (high: R2 = 0.97, P < 0.01; moderate: R2 = 0.93, P < 0.01). During exercise, slope ( P = 0.49) and y-intercept ( P = 0.91) did not differ between intensities. In contrast, slope was steeper ( P < 0.01) and y-intercept was higher ( P < 0.01) during recovery from high-intensity exercise. Thermal behavior is greater during high-intensity exercise because of the greater stimulus to behave. The withdrawal of thermal behavior is augmented after high-intensity exercise. NEW & NOTEWORTHY This is the first study to determine the effects of exercise intensity on thermal behavior. We show that exercise intensity does not independently modulate thermal behavior during exercise but is dependent on the magnitude of afferent stimuli. In contrast, the withdrawal of thermal behavior after high-intensity exercise is augmented. This may be a consequence of an attenuated perceptual response to afferent stimuli, which may be due to processes underlying postexercise hypoalgesia.
36

Parmenter, M. A., M. M. Manore und J. T. Daniels. „EPOC FOLLOWING HIGH INTENSITY INTERMITTENT AND MODERATE INTENSITY CONTINUOUS EXERCISE“. Medicine & Science in Sports & Exercise 33, Nr. 5 (Mai 2001): S73. http://dx.doi.org/10.1097/00005768-200105001-00416.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
37

Boutcher, Stephen H. „High-Intensity Intermittent Exercise and Fat Loss“. Journal of Obesity 2011 (2011): 1–10. http://dx.doi.org/10.1155/2011/868305.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
The effect of regular aerobic exercise on body fat is negligible; however, other forms of exercise may have a greater impact on body composition. For example, emerging research examining high-intensity intermittent exercise (HIIE) indicates that it may be more effective at reducing subcutaneous and abdominal body fat than other types of exercise. The mechanisms underlying the fat reduction induced by HIIE, however, are undetermined. Regular HIIE has been shown to significantly increase both aerobic and anaerobic fitness. HIIE also significantly lowers insulin resistance and results in a number of skeletal muscle adaptations that result in enhanced skeletal muscle fat oxidation and improved glucose tolerance. This review summarizes the results of HIIE studies on fat loss, fitness, insulin resistance, and skeletal muscle. Possible mechanisms underlying HIIE-induced fat loss and implications for the use of HIIE in the treatment and prevention of obesity are also discussed.
38

Byrd, S. K., L. J. McCutcheon, D. R. Hodgson und P. D. Gollnick. „Altered sarcoplasmic reticulum function after high-intensity exercise“. Journal of Applied Physiology 67, Nr. 5 (01.11.1989): 2072–77. http://dx.doi.org/10.1152/jappl.1989.67.5.2072.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
This study examined the effects of acute high-intensity exercise on the rate and capacity of Ca2+ uptake and Ca2+-stimulated adenosinetriphosphatase (ATPase) activity of the sarcoplasmic reticulum and the reversibility of these effects. Thoroughbred horses were run at maximal O2 uptake on a high-speed treadmill until fatigued. Muscle temperatures and biopsy samples were collected at rest, immediately after exercise, and 30 and 60 min after exercise. Blood samples were collected at rest and 5 min after exercise. Muscle and blood (lactate concentration) were three- and fivefold greater than pre-exercise values. Muscle temperature and pH immediately after post-exercise were 43 degrees C and 6.55, respectively, but approached rest values by 60 min after exercise. The initial rate and maximal capacity of Ca2+ uptake of muscle homogenates and isolated sarcoplasmic reticulum were significantly depressed immediately after exercise. This depression was paralleled by decreased activity of the Ca2+-stimulated ATPase. However, both Ca2+ uptake (rate and capacity) and Ca2+4-ATPase activity had returned to normal by 60 min after exercise. These findings demonstrate that changes in sarcoplasmic reticulum function after high-intensity exercise may be induced but not sustained by local changes in muscle pH and/or temperature.
39

Marcora, Samuele Maria, und Walter Staiano. „Reply to: What limits exercise during high-intensity aerobic exercise?“ European Journal of Applied Physiology 110, Nr. 3 (02.07.2010): 663–64. http://dx.doi.org/10.1007/s00421-010-1563-y.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
40

Hellsten, Y., B. Sjödin, E. A. Richter und J. Bangsbo. „Urate uptake and lowered ATP levels in human muscle after high-intensity intermittent exercise“. American Journal of Physiology-Endocrinology and Metabolism 274, Nr. 4 (01.04.1998): E600—E606. http://dx.doi.org/10.1152/ajpendo.1998.274.4.e600.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
The exchange of purines in exercised and rested muscle and their relation to muscle ATP levels after intense intermittent exercise were investigated. Seven subjects performed one-legged knee extensor exercise on the following two occasions: without (control; C) and with (high purines; HP) additional arm exercise. There was a greater net release of hypoxanthine by the exercised muscle during the recovery period in HP compared with C [185 ± 44 vs. 101 ± 30 (SE) μmol/kg muscle; P < 0.05]. During recovery, the arterial urate concentration was higher in HP compared with C (peak: 585 ± 48 vs. 355 ± 20 μmol/l; P < 0.05). The exercised but not the rested muscle extracted a marked amount of urate (330 μmol/kg muscle) from plasma in the HP trial. Muscle ATP levels after 90 min of recovery in HP were lower than at rest (24.3 ± 0.6 vs. 20.1 ± 1.1 mmol/kg dry wt). The present data suggest that a single session of long-term high-intensity intermittent exercise causes a significant release of purines from the muscle into blood, which contributes to a sustained lowered level of the muscle ATP concentration. Furthermore, intensely exercised muscle extracts urate when plasma urate is elevated, an event that may be of importance for the replenishment of oxidized muscle urate stores.
41

Padilha, Camila S., Francois Billaut, Caique Figueiredo, Valéria Leme Gonçalves Panissa, Fabrício Eduardo Rossi und Fabio S. Lira. „Capsaicin Supplementation during High-intensity Continuous Exercise: A Double-blind Study“. International Journal of Sports Medicine 41, Nr. 14 (21.07.2020): 1061–66. http://dx.doi.org/10.1055/a-1088-5388.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
AbstractTo investigate the effect of acute capsaicin (CAP) supplementation on time to exhaustion, physiological responses and energy systems contribution during continuous high-intensity exercise session in runners. Fifteen recreationally-trained runners completed two randomized, double-blind continuous high-intensity exercises at the speed eliciting 90% V̇O2peak (90% s V̇O2peak), 45 minutes after consuming capsaicin or an isocaloric placebo. Time to exhaustion, blood lactate concentration, oxygen consumption during and 20-min post-exercise, energy systems contribution, time to reach V̇O2peak, heart rate and the rate of perceived exertion (RPE) were evaluated. There was no significant difference between conditions for time to reach V̇O2peak (CAP:391.71±221.8 vs. PLA:298.20±174.5 sec, ES:0.58, p=0.872), peak lactate (CAP:7.98±2.11 vs. PLA:8.58±2.15 µmol, ES:−0.28, p=0.257), time to exhaustion (CAP:654.28±195.44 vs. PLA:709.20±208.44 sec, ES:−0.28, p=0.462, end-of-exercise heart rate (CAP:177.6±14.9 vs. PLA:177.5±17.9 bpm, ES:−0.10, p=0.979) and end-of-exercise RPE (CAP: 19±0.8 vs. PLA: 18±2.4, ES: 0.89, p=0.623). In conclusion, acute CAP supplementation did not increase time to exhaustion during high-intensity continuous exercise nor alter physiological responses in runners.
42

A. Porto, Andrey, Vin韈ius Koti Kamada, Leticia Santana Oliveira, Joana Z. Chambrone, David M. Garner und Vitor E. Valenti. „Cardiac Autonomic Modulation and High–Intensity Exercise“. Journal of Cardiology and Therapy 3, Nr. 3 (2016): 515–18. http://dx.doi.org/10.17554/j.issn.2309-6861.2016.03.101-1.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
43

CHIA, Michael. „Aerobic Energy Contribution during High Intensity Exercise“. Asian Journal of Physical Education & Recreation 10, Nr. 2 (01.12.2004): 15–21. http://dx.doi.org/10.24112/ajper.101143.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
Annotation:
LANGUAGE NOTE | Document text in English; abstract also in Chinese.The review focuses on the aerobic energy contribution during high intensity cycling exercise. It is erroneous to assume that the energy demands of an exercise task can be met exclusively by either aerobic or anaerobic sources. During peak oxygen uptake determination, especially during the latter portions of the incremental exercise test, the anaerobic energy stores are also taxed. Not surprising, during maximal exercise of a short duration, there is also energy supplementation from aerobic energy sources. However, for a test to be considered predominantly anaerobic, the aerobic contribution to the test must be kept minimal. The quantification of aerobic contribution to a maximal exercise performance is difficult because the mechanical efficiency (ME) during a non-steady state exercise task remains speculative. Nevertheless extreme ME values for cycling have been proposed to provide a general scope of the estimated values. In adults, assumptions about oxygen uptake lag time, the size and role of the stored oxygen stores, which are taken into account also affect the magnitude to the aerobic contribution. Equivalent data on young people are insecure and greater research attention in this area is advised.本文著重介紹了大強度自行車運動中的有氧供能。如果認為運動中機體所需能量僅以某一能源系統,有氧系統或無氧系統供能是不正確的。在逐級遞增負荷測定最大攝氧量的運動中,尤其在測試的後階段,無氧系統參與供能。而在短時間的最大強度運動中,有氧供能也佔有一定的比例。即使進行無氧運動,在測試中仍能發現有低比例有氧供能。很難確定有氧系統在最大強度運動中的供能量為多少,因為不穩定狀態下的運動其供能效率仍不十分明確。但對於踏車運動中最高供能效率有一估計值範圍。對於成年人,攝氧量的延遲時間以及氧的儲存量的多少將影響最大有氧供能的比例。而在青少年中:有關這方面的資料較為缺乏,有待進一步的研究。
44

Hargreaves, Mark. „Molecular Signalling during High Intensity Exercise Training“. Medicine & Science in Sports & Exercise 39, Supplement (Mai 2007): 52. http://dx.doi.org/10.1249/01.mss.0000272526.63676.2b.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
45

Medbø, J. I., und O. M. Sejersted. „Plasma potassium changes with high intensity exercise.“ Journal of Physiology 421, Nr. 1 (01.02.1990): 105–22. http://dx.doi.org/10.1113/jphysiol.1990.sp017935.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
46

Peterson, Margaret G. E., Sandy B. Ganz, John P. Allegrante und Charles N. Cornell. „High-Intensity Exercise Training Following Hip Fracture“. Topics in Geriatric Rehabilitation 20, Nr. 4 (Oktober 2004): 273–84. http://dx.doi.org/10.1097/00013614-200410000-00007.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
47

MOON, MARY ANN. „High-Intensity ‘Exergames’ Motivate Children to Exercise“. Family Practice News 41, Nr. 7 (April 2011): 44. http://dx.doi.org/10.1016/s0300-7073(11)70361-9.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
48

Sarmiento, Samuel, Juan Manuel García-Manso, Juan Manuel Martín-González, Diana Vaamonde, Javier Calderón und Marzo Edir Da Silva-Grigoletto. „Heart rate variability during high-intensity exercise“. Journal of Systems Science and Complexity 26, Nr. 1 (Februar 2013): 104–16. http://dx.doi.org/10.1007/s11424-013-2287-y.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
49

Burns, Stephen F., Masashi Miyashita und David J. Stensel. „High-Intensity Interval Exercise and Postprandial Triacylglycerol“. Sports Medicine 45, Nr. 7 (08.04.2015): 957–68. http://dx.doi.org/10.1007/s40279-015-0327-6.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen
50

Mookerjee, Swapan, Ansgar Steegmans, Uwe Drescher, Axel Knicker und Uwe Hoffmann. „Hemodynamic Responses Following High Intensity Isokinetic Exercise“. Medicine & Science in Sports & Exercise 42 (Oktober 2010): 79–80. http://dx.doi.org/10.1249/01.mss.0000389412.11569.25.

Der volle Inhalt der Quelle
APA, Harvard, Vancouver, ISO und andere Zitierweisen

Zur Bibliographie