首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Summary The effects of intravenous injections of Atropine (1.8 mg) and practolol (15 mg) on the thermoregulatory responses to 1 h of exercise on a motordriven treadmill have been investigated on six healthy subjects.The results show that -blockade had little effect on thermal responses to work except for a small but significant (p<0.05) decrease in mean skin temperature (¯T sk ) and peripheral tissue heat conductance (K). Metabolic (M) and total heat (H) production, and evaporative sweat loss (E) and rectal temperature (T re ) were similar to control values. In contrast, atropine, particularly at work loads beyond 60% maximal aerobic power output (VO2 max), raised T re (p<0.001), ¯T sk (p<0.001) and reduced E by approximately 50%. At the highest work loads T re increased as a linear function of time during the latter part of exercise, and at the 60th min was almost independent of relative stress (expressed as % VO2 max) imposed on the subjects. At the lower work loads the majority of subjects reached thermal equilibrium before the end of exercise by maintaining their convective heat transfer from core to periphery by increasing peripheral blood flow (as indicated by K), and raising their heat losses to environment by convection and radiation. The latter pathways for heat dissipation were enhanced by the subjects ability to sustain a ¯T sk 4 C above control values independently of M. Atropine had no effect on M or H but greatly affected work performance, no subject was able to exercise at loads >70% VO2 max for 1 h. These results demonstrate the ability of the thermoregulatory system to adapt to -adrenergic and to parasympathetic blockade during light exercise, and underline the effects of a reduction in the capacity of the sweating mechanism on physiological performance at higher rates of work.List of Abbreviations used in the Text M Metabolic heat production - H Total heat production - E Evaporative sweat loss - T re Rectal temperature - ¯T sk Mean skin temperature - K Peripheral tissue heat conductance - PBF Peripheral blood flow - VO2 max Maximal aerobic power output - f H Cardiac frequency  相似文献   

2.
Early studies have demonstrated that rectal temperature (T re) decreases and mean skin temperature (T sk) increases in subjects changing their posture from standing to supine, and vice versa. Such changes have important implications insofar as thermal stress experiments are conducted and interpreted. However, the extent of these changes between steady-state conditions is not known. In addition, it is not known whether thermal balance is also affected by postural changes. To examine these questions, 11 healthy males were exposed to a thermoneutral air environment (28.2–28.5°C and 40% relative humidity) in various postures at rest. Body temperatures, heat losses, and metabolic rate were measured. Subjects wore shorts only and began in an upright posture (standing or sitting at an inclination of 7.5°) on a customized tilt-table. They were tilted twice, once into a supine position and then back to the original upright position. Each tilt occurred after steady state was satisfied based on the subject's circadian variation of T re determined previously in a 4.25 h control supine trial. Times to supine steady state following the first tilt were [mean (SE)] 92.6 (6.4) and 116.6 (5.1) min for the standing and sitting trials, respectively. Times to upright steady state following the second tilt were 107.9 (11.4) and 124.1 (9.0) min. Mean steady-state T re and T sk were 36.87 (0.07) and 34.04 (0.14), 37.47 (0.09) and 33.48 (0.14), and 37.26 (0.05) and 33.49 (0.10) °C for supine, standing, and sitting, respectively. Thermal balance was attained in all steady-state conditions, and allowing for a decrease in the weighting factor of T re for mean body temperature in the upright postures, it also appears that thermal balance was preserved between changes in posture. These results are consistent with no perceived changes by the subjects in their thermal comfort and skin wetness.  相似文献   

3.
Body heat storage can be estimated by calorimetry (from heat gains and losses) or by thermometry [from changes (Δ) in mean body temperature (T b) calculated as a weighted combination of rectal (T re) and mean skin temperatures (T sk)]. If an invariant weighting factor ofT re andT sk were to be used (for instance, ΔT b = 0.8 · ΔT re + 0.2 · ΔT sk under hot conditions), body heat storage could be over- or underestimated substantially relative to calorimetry, depending on whether the subject was wearing light or protective clothing. This study investigated whether discrepancies between calorimetry and thermometry arise from methodological errors in the calorimetric estimate of heat storage, from inappropriate weightings in the thermometric estimate, or from both. Residuals of calorimetry versus thermometric estimates were plotted against individual variables in the standard heat balance equation, applying various weighting factors toT re andT sk. Whether light or protective clothing was worn, the calorimetric approach generally gave appropriate estimates of heat exchange components and thus heat storage. One exception was in estimating latent heat loss from sweat evaporation. If sweat evaporation exceeded 650 g·h−1 when wearing normal clothing, evaporative heat loss was overestimated and thus body heat storage was underestimated. Nevertheless, if data beyond this ceiling were excluded from the analyses, the standard 4:1 weighting matched calorimetric heat storage estimates quite well. When wearing protective clothing, the same 4:1 weighting approximated calorimetric heat storage with errors of less than approximately 10%, but only if environmental conditions allowed a subject to exercise for more than 90 min. The best thermometric estimates of heat storage were provided by using two sets of relative weightings, based upon the individual's metabolic heat production ( in kilojoules per metre squared per hour): {4 − [( )· ] 2}:1 for an initial, thermoneutral environment and {4 + [( ) · ] · 5}: 1 for a final, hot environment; the optimal value of lay between 450 and 500 kJ m−2 · h−1. We concluded that the accuracy of thermometric estimates of heat storage can be improved by modifying weighting factors ofT re andT sk according to the environment, type of clothing, and metabolic rate.  相似文献   

4.
To determine the effects of pre-warming on the human metabolic and thermoregulatory responses to prolonged steady-rate exercise in moderate ambient temperatures and relative humidities [means (SD) 21.7 (2.1)° C and 36.7 (5.4)%, respectively], six healthy men each ran at a steady-rate (70% maximal oxygen uptake) on a treadmill until exhausted after being actively pre-warmed (AH), passively pre-warmed (PH), and rested (Cont). Exercise time to exhaustion was significantly reduced following both AH and PH compared to Cont [AH 47.8 (14.0) min, PH 39.6 (16.0) min, Cont 62.0 (8.8) min; P<0.05]. During exercise there were no significant differences in oxygen uptake, total sweat loss, mean skin temperature (Tsk) and the thermal gradient (T re–Tsk, where T re is rectal temperature) following the three conditions. Serum prolactin, plasma catecholamine and plasma free fatty acid concentrations were also similar between all three trials. In contrast, T re, mean body temperature, heart rate and ratings of perceived exertion were significantly greater during the initial 25 min of exercise following both AH and PH, compared with Cont (P<0.05). At exhaustion, there were no significant differences in the metabolic and thermoregulatory responses to exercise between the trials. The current findings demonstrate that AH and PH promote a reduction in prolonged submaximal endurance performance under moderate environmental temperatures compared with pre-exercise rest. Such observations appear likely to have been mediated through mechanisms associated with the earlier development of high internal body temperature which resulted in changes in the capacity for heat storage. Electronic Publication  相似文献   

5.
The effect of ambient temperature on gross-efficiency in cycling   总被引:1,自引:0,他引:1  
Time-trial performance deteriorates in the heat. This might potentially be the result of a temperature-induced decrease in gross-efficiency (GE). The effect of high ambient temperature on GE during cycling will be studied, with the intent of determining if a heat-induced change in GE could account for the performance decrements in time trial exercise found in literature. Ten well-trained male cyclists performed 20-min cycle ergometer exercise at 60% (power output at which VO2max was attained) in a thermo-neutral climate (N) of 15.6 ± 0.3°C, 20.0 ± 10.3% RH and a hot climate (H) of 35.5 ± 0.5°C, 15.5 ± 3.2% RH. GE was calculated based on VO2 and RER. Skin temperature (T sk), rectal temperature (T re) and muscle temperature (T m) (only in H) were measured. GE was 0.9% lower in H compared to N (19.6 ± 1.1% vs. 20.5 ± 1.4%) (P < 0.05). T sk (33.4 ± 0.6°C vs. 27.7 ± 0.7°C) and T re (37.4 ± 0.6°C vs. 37.0 ± 0.6°C) were significantly higher in H. T m was 38.7 ± 1.1°C in H. GE was lower in heat. T m was not high enough to make mitochondrial leakage a likely explanation for the observed reduced GE. Neither was the increased T re. Increased skin blood flow might have had a stealing effect on muscular blood flow, and thus impacted GE. Cycling model simulations showed, that the decrease in GE could account for half of the performance decrement. GE decreased in heat to a degree that could explain at least part of the well-established performance decrements in the heat.  相似文献   

6.
Summary The thermoregulatory responses to 1 h exercise of 14 male (age range 18–65 year) and 7 female (age range 18–46 year) athletes and 4 (3 and 1 ) non-athletic subjects have been investigated in a moderate environment (T db=21 C, T wb=15 C and rh<50%) and analysed in relation to age, sex, and maximum aerobic power output (VO2 max).The maximal sweat loss (M sw max) under the given conditions was closely related (r=+0.90) to VO2 max and for a given relative work load (%VO2 max), rectal (T re) and mean skin (¯T sk) temperatures was the same in all subjects.Sweat loss (004d sw) was linearly related to total heat production (H) and to peripheral tissue heat conductance (K) and if expressed in relative terms (%M sw max) was linearly related to T re. For a given T re relative sweat rate was identical in the groups studied. From these results it would seem that during exercise T re rises to meet the requirements of heat dissipation by establishing a thermal gradient from core to skin and stimulating sweating in proportion to maximal capacity of the system. Thus provided the thermal responses to work were standardised using the appropriate physiological variables, there was no evidence to be found for differences in thermoregulatory function which could be ascribed to sex or age.  相似文献   

7.
This study determined whether a torso-vest forced ambient air body ventilation system (BVS) reduced physiological strain during exercise-heat stress. Seven heat-acclimated volunteers attempted nine, 2-h treadmill walks at 200 W m−2 in three environments, −40°C, 20% rh (HD), 35°C, 75% rh (HW), and 30°C, 50% rh, (WW) wearing the Army Combat Uniform, interceptor body armor (IBA) and Kevlar helmet. Three trials in each environment were BVS turned on (BVSOn), BVS turned off (BVSOff), and no BVS (IBA). In HD, BVSOn significantly lowered core temperature (T re), heart rate (HR), mean skin temperature (T sk), mean torso skin temperature (T torso), thermal sensation (TS), heat storage (S), and physiological strain index (PSI), versus BVSOff and IBA (P < 0.05). For HW (n = 6), analyses were possible only through 60 min. Exercise tolerance time (min) during HW was significantly longer for BVSOn (116 ± 10 min) versus BVSOff (95 ± 22 min) and IBA (96 ± 18 min) (P < 0.05). During HW, BVSOn lowered HR at 60 min versus IBA, T sk from 30 to 60 min versus BVSOff and IBA, and PSI from 45 to 60 min versus BVSOff and at 60 min versus IBA (P < 0.05). BVSOn changes in T re and HR were lower in HD and HW. During WW, BVSOn significantly lowered HR, T sk, and T torso versus BVSOff and IBA (P < 0.05) during late exercise. Sweating rates were significantly lower for BVSOn versus BVSOff and IBA in both HD and WW (P < 0.05), but not HW. These results indicate that BVSOn reduces physiological strain in all three environments by a similar amount; however, in hot-dry conditions the BVSOff increases physiological strain.  相似文献   

8.
Summary Well matched unacclimatised older (age 55–68, 4 women, 2 men) and younger (age 19–30, 4 women, 2 men) subjects performed 75 min cycle exercise (40% ) in a hot environment (37°C, 60% rh). Rectal temperature (T re), mean skin temperature (¯T sk), arm blood flow (ABF, strain gauge plethysmography), and cardiac output (Q, CO2 rebreathing) were measured to examine age-related differences in heat-induced vasodilatation.T re and¯T sk rose to the same extent in each group during the exposure. There was no significant intergroup difference in sweat rate (older: 332±43 ml · m–2 · h–1, younger: 435±49 ml · m–2 · h–1; mean±SEM). However, the older subjects responded to exercise in the heat with a lower ABF response which could be attributed to a lower for the same exercise intensity. The slope of the ABF-T re relationship was attenuated in the older subjects (9.3±1.3 vs 17.9±3.3 ml · 100 ml–1 · min–1 · °C–1,p <0.05), but theT re threshold for vasodilatation was about 37.0°C for both groups. These results suggest an altered control of skin vasodilatation during exercise in the heat in older individuals. This attenuated ABF response appears to be unrelated to , and may reflect an age-related change in thermoregulatory cardiovascular function.  相似文献   

9.
The study was conducted to investigate the thermoregulation of young children compared to that of adults. A group of 19 children (ages 9 months-4.5 years), with only 3 children aged 3 years or above, and 16 adults first rested in a thermoneutral room (air temperature 25°C relative humidity 50%, air velocity 0.2 m·s–1). They were then exposed to a hot room (air temperature 35°C, relative humidity 70%, air velocity 0.3 m·s) next door for 30 min, and then returned to the thermoneutral room where they stayed for a further 30 min. The rectal temperature (T re), skin temperatures (T sk) at seven sites, heart rate (HR), total sweat rate ( ), local sweat rate ( ) and the Na+ concentration of the sweat were measured. There was no significant difference inT re between the children and their mothers in the rest phase. However, theT re of the children increased as soon as they entered the hot room and was significantly higher than during the control period, and than that of the mothers during heat exposure. MeanT sk, forehead, abdomen and instepT sk were significantly higher in the children during both the thermoneutral and heat exposure. The was significantly higher and Na+ concentrations in the sweat on the back and upperarm were significantly lower for the children during the heat exposure. They had a greater body surface area-to-mass ratio than the mothers by 64%, which indicated that they had advantages for thermal regulation. However, the sweating andT sk responses of the children were not enough to prevent a rise in body temperature. These results would suggest that the young children had the disadvantage of heating up easily due to their smaller body sizes and there may be maturation-related differences in thermoregulation during the heat exposure between young children and mothers.  相似文献   

10.

Purpose

To assess heat balance status of newborn infants nursed under radiant warmers (RWs) during intensive care.

Methods

Heat balance, thermal status and primary indicators of physiological strain were concurrently measured in 14 newborns nursed under RWs for 105 min. Metabolic heat production (M), evaporative heat loss (E), convective (C) and conductive heat flow (K), rectal temperature (T re) and mean skin temperatures (T sk) were measured continuously. The rate of radiant heat required for heat balance (R req) and the rate of radiant heat provided (R prov) were derived. The rate of body heat storage (S) was calculated using a two-compartment model of ‘core’ (T re) and ‘shell’ (T sk) temperatures.

Results

Mean M, E, C and K were 10.5 ± 2.7 W, 5.8 ± 1.1 W, 6.2 ± 0.8 W and 0.1 ± 0.1 W, respectively. Mean R prov (1.7 ± 2.6 W) and R req (1.7 ± 2.7 W) were similar (p > 0.05). However, while the resultant mean change in body heat content after 105 min was negligible (–0.1 ± 3.7 kJ), acute time-dependent changes in S were evidenced by a mean positive heat storage component of +6.4 ± 2.6 kJ and a mean negative heat storage component of –6.5 ± 3.7 kJ. Accordingly, large fluctuations in both T re and T sk occurred that were actively induced by changes in RW output. Nonetheless, no active physiological responses (heart rate, breathing frequency and mean arterial pressure) to these bouts of heating and cooling were observed.

Conclusions

RWs maintain net heat balance over a prolonged period, but actively induce acute bouts of heat imbalance that cause rapid changes in T re and T sk. Transient bouts of heat storage do not exacerbate physiological strain, but could in the longer term.  相似文献   

11.
We evaluated the cooling rate of hyperthermic subjects, as measured by three estimates of deep core temperatures (esophageal, rectal and aural canal temperatures), during immersion in a range of water temperatures. The objective of the study was to compare the three indices of core temperature and define safe cooling limits when using rectal temperature to avoid the development of hypothermia. On 4 separate days, seven subjects (four males, three females) exercised for 45.4±4.1 min at 65% at an ambient temperature of 39°C, RH: 36.5%, until rectal temperature (T re) increased to 40.0°C (39.5°C for two subjects). Following exercise, the subjects were immersed in a circulated water bath controlled at 2, 8, 14 and 20°C until T re returned to 37.5°C. When T re reached normothermia during the cooling period (37.5±0.05°C), both esophageal (T es) (35.6±1.3°C) and aural canal (T ac) (35.9±0.9°C) temperatures were approaching or reaching hypothermia, particularly during immersion in 2°C water (T es=34.5±1.2°C). On the basis of the heat loss data, the heat gained during the exercise was fully eliminated after 5.4±1.5, 7.9±2.9, 10.4±3.8 and 13.1±2.8 min of immersion in 2, 8, 14 and 20°C water, respectively, with the coldest water showing a significantly faster cooling rate. During the immersion in 2°C water, a decrease of only 1.5°C in T re resulted in the elimination of 100% of the heat gained during exercise without causing hypothermia. This study would therefore support cooling the core temperature of hyperthermic subjects to a rectal temperature between 37.8°C (during immersion in water >10°C) and 38.6°C (during immersion in water <10°C) to eliminate the heat gained during exercise without causing hypothermia.  相似文献   

12.
Summary Eight young, sedentary men (aged 34 years, SD 3) and six older moderately active, unacclimated men (aged 57 years, SD 2) walked on a treadmill at 30% of their maximum oxygen consumption up to 3.5 h in a thermoneutral [dry bulb temperature (T db) 21°C, relative humidity (r.h.) 43%)], a warm humid (T db 30°C, r.h. 80%) and a hot dry (T db 40°C, r.h. 20%) environment while wearing ordinary working clothes (0.7 c/o). Their oxgen consumption, heart rate (f c), rectal (T re) and mean skin temperature (Tsk), sweat rate (SR), and evaporative rate (ER) were measured during the tests. The ratings of thermal sensation (TS) and perceived exertion (RPE) were assessed using standard scales. In the heat stress tests, the number of experiments discontinued did not significantly differ between the two groups. The mean levels and end-exercise values of T re, Tsk, f c, TS and RPE were not significantly different between the young and older subjects in any of the environments. In the warm humid environment, however, the T re and RPE of the older subjects increased continuously (P<0.05) during the test compared to the young subjects. No significant difference between the groups was observed in SR or in ER. In the hot dry environment, however, the ER of older men increased more slowly compared to the young men. In spite of some time-related differences observed in T re, RPE, and ER, the older subjects did not exhibit higher f c during exercise in the heat, they were not more hyperthermic and their performance times were similar to the young subjects. Therefore, it was concluded that older calendar age is not necessarily associated with a reduced ability to exercise in a hot environment and other factors, such as physical activity habits and aerobic capacity, may be equally important in determining heat tolerance in the elderly.  相似文献   

13.
The response of core temperature to exercise was investigated during recovery in order to avoid the antagonistic competition between exercise and thermal reflexes for the same effector systems which control skin blood flow. Five healthy, non-training males [mean (SD) age, 23.8 (2.04) years] were habituated to 29° C at relative 50% humidity for more than 2 h and then exercised by treadmill running at about 75% maximum oxygen uptake for 18 min. They then remained at 29° C for up to 65 min of recovery. Oesophageal (T es), rectal (T re) and skin temperatures (T sk) were recorded at 5-s intervals throughout. The abrupt fall of temperature gradient from the forearm to finger was used to identify the T es for skin vessel dilatation (T dil) during exercise. Mean (SE) Ts rose from a resting value of 36.67 (0.15)° C to 38.22 (0.24)° C, mean T re rose from 37.09 (0.25)° C to 38.23 (0.15)° C, and T dil occurred at 37.39 (0.32)° C. Within 10 min of recovery mean T es fell to 37.31 (0.24)° C, where it remained a significant 0.64° C above its pre-exercise (PrEx) level (P0.018) but insignificantly different from T dil for the remaining 55 min of recovery. Meanwhile, T re fell gradually throughout recovery to 37.64 (0.18)° C. The T sk at all non-acral sites except the thigh had recovered to PrEx levels by 20–30 min post-exercise (PoEx). The rapid PoEx fall of T es to the level of T dil and the subsequent plateau above PrEx values suggests that heat dissipation during recovery was primarily passive once T es had fallen to T dil, even though T es and T re were significantly elevated. The relationship of these results to the set-point and load error concepts of thermal control is discussed.These data have been presented at the Canadian Physiological Society Winter meeting, January 1993, but have not been previously published  相似文献   

14.
Thermoregulatory responses of young and older men to cold exposure   总被引:2,自引:0,他引:2  
Summary Nine young (20–25 years) and ten older (60–71 years) men, matched for body fatness and surface area :mass ratio, underwent cold tests in summer and winter. The cold tests consisted of a 60-min exposure, wearing only swimming trunks, to an air temperature of 17°C (both seasons) and 12°C (winter only). Rectal (T re) and mean skin ( sk) temperatures, metabolic heat production (M), systolic (BPS) and diastolic (BPd) blood pressures and heart rate (f c) were measured. During the equilibrium period (28°C air temperature) there were no age-related differences inT re, sk, BPS, BPd, orf c regardless of season, although M of the older men was significantly lower (P<0.003). The decrease inT re and sk (due to the marked decrease in six of the older men) and the increase in BPS and BPd were significantly greater (P<0.004) for the older men during all the cold exposures. The rate of increase inM was significantly greater (P<0.01) for the older group when exposed to 12°C in winter and 17°C in summer (due to the marked increase in four of the older men). This trend was not apparent during the 17°C exposure in winter. There was no age-related difference in fc during the exposures. Significant decreases inT re and sk and increases inM, BPS and BPd during the 12°C exposure were observed for the older group (P< 0.003) compared to their responses during the 17°C exposure in winter. In contrast,T re,M, BPS in the young group were not affected as much by the colder environment. It was concluded that older men have more variable responses and some appear more or less responsive to mild and moderate cold air than young men.  相似文献   

15.
This study examined the effectiveness of endurance training and heat acclimation in reducing the physiological strain imposed by exercising in the heat while wearing protective clothing. Seven young men underwent 8 weeks of physical training [60–80% maximal aerobic power (VO2max) for 30–45 min · day–1, 3–4 days · week–1 at < 25° C] followed by 6 days of heat acclimation (45–55% VO2max for 60 min · day–1 at 40° C, 30% relative humidity). Nine other young men underwent corresponding periods of control observation and heat acclimation. Before and after each treatment, subjects completed a treadmill walk (4.8 km · h–1, 2% grade) in a climatic chamber (40° C, 30% relative humidity), wearing in turn normal combat clothing or clothing protecting against nuclear, biological, and chemical (NBC) agents. Criteria for halting this test were: (1) a rectal temperature (T re) of 39.3° C; (2) a heart rate (f c) 95% of the subject's observed maximum, maintained for 3 min; (3) unwillingness of the subject to continue; (4) the elapse of 120 min. The training regimen increased mean VO2max by 16% and mean plasma volume by 8%. When tested in normal combat clothing, the rates of increase in T re and f c were slower after training. However, when wearing NBC protective clothing, the only significant change induced by training was a higher mean skin temperature (T sk) in the early part of the test. Heat acclimation increased the mean plasma volume of untrained subjects by 8%, but their VO2max remained unchanged. When tested in normal combat clothing, acclimation decreased their mean values of T re, T sk, f c, and metabolic rate. When wearing NBC protective clothing, the only significant decrease after acclimation was in overall T re. In trained subjects, heat acclimation induced no further improvement in any physiological variable when wearing normal combat clothing, but reduced overall T re and T sk when wearing NBC protective clothing. Training- or acclimation-induced increases of sweat secretion (an average increment of 0.14–0.23 kg · h–1) were not accompanied by any statistically significant increase in sweat evaporation when wearing NBC protective clothing. Moreover, tolerance times were unchanged in either normal combat (116–120 min) or NBC protective clothing (47–52 min). We conclude that neither endurance training nor heat acclimation do much to improve exercise tolerance when wearing NBC protective clothing in hot environments, because any added sweat secretion decreases blood volume and increases discomfort without augmenting body cooling.  相似文献   

16.
Many thermal physiologists follow the conventional wisdom that physiological heat adaptations occur in the summer for people living in a humid continental climate (e.g. Central Canada, North-eastern and Mid-western United States and Eastern Europe); therefore experimentation across seasons is often avoided. However, since modern behavioral adaptations, such as air conditioning, are accessible and commonplace, it is not clear whether such physiological adjustments actually do occur. It was hypothesized that despite warm weather, residing in a humid continental climate throughout a summer will not elicit any significant physiological heat adaptations since the environmental stimulus for such adjustments will be mitigated by behavioral adaptations. Eight young healthy male volunteers cycled at 60% VO2max for 90-min in a temperate environment before (mid-May) and at the end of (start of September) summer. Core temperature [measured in the esophagus (T es), rectum (T re) and aural canal (T au)], mean skin temperature (T sk), forearm skin blood flow (SkBf), upper back sweat rate (LSR) and heart rate (HR) were measured throughout exercise. Weekly activity logs and a lifestyle questionnaire were also administered throughout the summer months. No significant differences between pre- and end-summer were observed throughout exercise for T es (p = 0.565), T re (p = 0.350), T au (p = 0.261), T sk (p = 0.955), SkBf (p = 0.112), LSR (p = 0.394) or HR (p = 0.343). Likewise, the thermosensitivity and T es at the onset threshold for LSR (p = 0.177, p = 0.512) and SkBf (p = 0.805, p = 0.556) were also not significantly different. The apparent lack of heat acclimatization could be due to frequent air-conditioning use and an avoidance of outdoor activity during the hottest times of day but may also be due to a lack of environmental stimulus.  相似文献   

17.
Summary The thermal responses of two healthy male subjects have been studied at the same mean skin temperature (T sk ) during negative work, positive work and positive work in which additional heating was induced by diathermy. The results showed that for a given metabolic heat production (M) rectal (T re ) and oesophageal (T oes ) temperatures were higher in negative work and positive work with diathermy than normal control experiments. In resting experiments with diathermy, T oes rose to the same level as when an equal amount of heat was produced metabolically by exercise. In negative work and positive work with diathermy sweat loss (M sw ) was higher for a given M and T sk than found for normal exercise, but in all three forms of work the relationship of M sw to total heat production (H) was identical. During positive work with and without diathermy the differences in M sw could be accounted for by using a previously developed model of relative sweating rate: %M sw = – constant + T re (or T oes ) + T sk .In negative work, removal of the difference between predicted and observed %M sw required the inclusion of a further factor into the equation based on muscle temperature. The results suggest that the core temperature in exercise rises to meet the requirements of heat dissipation mainly by stimulating M sw and establishing a heat transfer gradient from core to periphery and is not necessarily or uniquely related to M or to the rate of working. The study underlines the usefulness of negative work and diathermy as physiological tools for the further understanding of thermoregulation during exercise.  相似文献   

18.
Eight older (60–65 years) and six younger (20–25 years) men were exposed to a standard heat stress for 60 min in summer, autumn, winter, and spring. The test consisted of placing the lower legs and feet in a 42°C water bath while sitting in constant environmental conditions (30°C and 45% relative humidity). The increase of rectal temperature (T re) was significantly greater (P < 0.05) in autumn, winter, and spring than in summer for the older group, but significantly greater only in winter than in summer for the younger group (P < 0.05). The T re was greater for the older group in all seasons, but of significance only in autumn and spring (P < 0.01). There were no significant season-related differences for metabolic heat production (m) and mean skin temperature ( sk) during the heat test in the respective groups, although the m and sk were lower for the older group in all seasons (P < 0.01). In the older group total body sweating rate (msw) divided by T re (total msw/T re) decreased from summer to winter (P < 0.02) and did not differ between winter and spring, whereas total msw/T re in the younger group increased in spring after decreasing from autumn to winter (P < 0.03). The variations of the value, local sweating rate on the back and thigh divided by T re (back msw/T re and thigh msw/T re), were similar to those of the total msw/T re in each group, except for back msw/T re in the younger group, which did not increase from winter to spring. The total msw/T re, back msw/T re and thigh msw/T re were significantly less for the older group in summer, autumn and spring (P < 0.05). The range of seasonal variations was significantly less for the older group (P < 0.001). The results indicated that, compared with younger men in older men, the enhancement of sweating function toward summer occurred later and its reduction toward winter occurred earlier despite a smaller range of seasonal variation and that older men had a somewhat lesser capability to maintainT re when challenged by heat stress in all seasons.  相似文献   

19.
Summary Comparisons of physiological responses to 0, 0.5, 1, and 2 mg atropine (IM) were made in seven males ( ± SD: age, 24±3 years; ht, 174±12 cm; wt, 76±3 kg) while they exercised (~ 390 W) in a hot-dry (40 C, 20% rh) environment. Responses to 4 mg, as well as repeatability of responses to 2 mg, were studied in two and six of these subjects, respectively. On 8 test days an intramuscular injection of atropine or saline control was administered 20 min before subjects walked on a treadmill for two 50-min bouts. Heart rate (HR) during exercise did not change in the control trial but by min 50 increased during all atropine trials (P<0.01). Rectal temperature Tre) increased (P<0.01) in all trials by min 50 and continued increasing (P<0.01) in the 2-mg trial during the second exercise bout. For the two subjects tested with all dosages (0.5–4 mg atropine), the change in HR and Tre between the atropine and control trials at 50 min of exercise was regressed against the various atropine dosages. The relationship (r=0.92) for HR was curvilinear while the relationship (r=0.99) for Tre was linear. Mean weighted skin temperature ( sk) was relatively constant during exercise and was warmer (P<0.05) with increasing atropine dosage. In a repeat 2 mg trial, HR was 6 bt·min–1 lower (P<0.05) on the second exposure but Tre was the same (P>0.05) on both days. For subjects walking in the heat, three new observations were: 1) 0.5 mg of atropine resulted in increased HR and sk compared to control values; 2) HR was elevated but the magnitude of change decreased with increasing dosage, while the elevation in Tre was consistent with increasing dosage; and 3) rectal temperatures (in trials with and without atropine) were unaffected by previous days of atropine administration.  相似文献   

20.
Summary Energy substrate mobilization has been suggested as being a limiting factor for the rate of cold-induced thermogenesis (M), and consequently in delaying hypothermia. The evidence supporting this hypothesis in humans, however, is not convincing and the hypothesis has yet to be tested in a rigorous manner using a full heat balance analysis (partitional calorimetry). The goal of this study was therefore to re-investigate whether enhancing energy substrate mobilization by feeding cold-exposed subjects would improveM and affect heat debt (S; the minute-by-minute balance ofM and heat losses) as well as rectal (T re) and mean skin temperatures . Nine healthy semi-nude fasted subjects were exposed to 5° C (3 h at rest, 1 m · s–1 wind) on three occasions following the ingestion at min 0 and 90 of either: (1) a placebo, (2) 710 kJ of pure carbohydrates (100%-CHO), or (3) 710 kJ of a high-carbohydrate bar (High-CHO). As expected in the cold,T re andT sk decreased whereasM, S and heat losses increased (P<0.01). However, there were no differences between treatments, including the finalT re [mean (SEM); 36.4 (0.2); 36.5 (0.3) and 36.5 (0.2)°C for the placebo, 100%-CHO and High-CHO tests, respectively]. During the 100%-CHO treatment, rates of carbohydrate oxidation were the highest and fat oxidation the lowest (P<0.05), whereas the High-CHO treatment caused smaller changes. The results demonstrate that in the cold, enhancing energy substrate mobilization by ingesting substrates in the form of a supplement containing either mainly or only CHO does not cause detectable changes inM, heat loss,S or body temperatures, compared to the ingestion of a placebo. Under the present conditions, the results do not support the theory that energy substrate mobilization is a limiting factor for cold-induced thermogenesis in humans.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号