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 15 毫秒
1.
A single 2h light pulse (250 lux) was given at various times to phase shift the locomotor circadian rhythm of two species of closely related cockroaches, Blattella bisignata and Blatella germanica. The phase-response curve (PRC) of both species showed a similar pattern. Phase delays and advances were induced by light pulse during the early and late subjective night, respectively, while no clear phase shifting was elicited during the subjective day. However, the magnitude of the phase delay (1.89h ± 0.66h) and advance (0.69h ± 0.36h) of B. bisignata was significantly larger than that of B. germanica (0.78h ± 0.38h and 0.35h ± 0.18h, respectively). This result indicates the superior adjustability of the circadian clock in B. bisignata. The periodresponse curve (PdRC) was also constructed for both species. Although both species did not show great flexibility in circadian period changes, the phase shifts were significantly correlated with the period changes in the advance zone of B. bisignata (r = 0.72, P <. 1). This allowed the circadian clock of B. bisignata to display better entrainability since the phase advance adjustment was significantly more difficult than that of phase delay. The results indicate the overall adjustability of the circadian clock of B. germanica is inferior to that of B. bisignata. The significance of this finding is discussed from an ecological perspective. (Chronobiology International, 18(5), 767– 780, 2001)  相似文献   

2.
The German cockroach, Blattella germanica, and the double-striped cockroach, B. bisignata, are sibling species with a similar period sequence but a distinctive circadian rhythm in locomotion. The cell distribution of immunoreactivity (ir) against three clock-related proteins, Period (PER), Pigment Dispersing Factor (PDF), and Corazonin (CRZ), was compared between the species. The PER-ir cells tend to form clusters and are sprayed out in the central nervous system. Three major PER-ir cells are located in the optic lobes, which are the sites of the major circadian clock. They are interconnected with PER-ir axon bundles. Interestingly, the potential output signal of the circadian clock, PDF, is co-localized with PER in all three groups of cells. However, only two CRZ-ir cells and their axons are found in the optic lobes and they are not co-localized with PER-ir or PDF-ir cells and axons. Since only one circadian rhythm is expressed in locomotion, the time signals from both major clocks in optic lobes are coupled by connection with PDF-ir axons. A group of 3-4 PER-ir cells in the protocerebrum display typical characteristics of neurosecretary cells. In addition, there are numerous, small PER-ir and PDF-ir co-localized cells in the pars intercerebralis (PI), which have direct connections with the neurohemoorgan, corpora cardiaca, through PER-ir and PDF-ir axons. Based on these findings, the cellular connection shows a circadian control through the endocrine route. For the rest of central nervous system, only a few PER-ir and PDF-ir cells or axons are detected. This finding implies the circadian clock for locomotion is not located in subesophageal ganglion, thoracic or abdominal ganglia, but may use other neural messengers to pass on circadian signals. Since the overall distribution pattern of the clock cells are the same for B. germanica and B. bisignata, the possible explanation for the different expressions of locomotion between the species depends on genes downstream of per, pdf, and crz.  相似文献   

3.
A musically enhanced bird song stimulus presented in the early subjective night phase delays human circadian rhythms. This study determined the phase-shifting effects of the same stimulus in the early subjective day. Eleven subjects (ages 18-63 yr; mean +/- SD: 28.0 +/- 16.6 yr) completed two 4-day laboratory sessions in constant dim light (<20 lux). They received two consecutive presentations of either a 2-h musically enhanced bird song or control stimulus from 0600 to 0800 on the second and third mornings while awake. The 4-day sessions employing either the stimulus or control were counterbalanced. Core body temperature (CBT) was collected throughout the study, and salivary melatonin was obtained every 30 min from 1900 to 2330 on the baseline and poststimulus/postcontrol nights. Dim light melatonin onset and CBT minimum circadian phase before and after stimulus or control presentation was assessed. The musically enhanced bird song stimulus produced significantly larger phase advances of the circadian melatonin (mean +/- SD: 0.87 +/- 0.36 vs. 0.24 +/- 0.22 h) and CBT (1.08 +/- 0.50 vs. 0.43 +/- 0.37 h) rhythms than the control. The stimulus also decreased fatigue and total mood disturbance, suggesting arousing effects. This study shows that a musically enhanced bird song stimulus presented during the early subjective day phase advances circadian rhythms. However, it remains unclear whether the phase shifts are due directly to effects of the stimulus on the clock or are arousal- or dim light-mediated effects. This nonphotic stimulus mediates circadian resynchronization in either the phase advance or delay direction.  相似文献   

4.
Light and serotonin were found to cause phase shifts of the circadian neural activity rhythm in the optic lobe of the cricket Gryllus bimaculatus cultured in vitro. The two phase-shifting agents yielded phase-response curves different in shape. Light induced phase delay and advance in the early and late subjective night, respectively, and almost no shifts in the subjective day, whereas serotonin phase-advances the clock during the subjective day and induced delay shifts during the subjective night. The largest phase advance and delay occurred at circadian time 21 and 12, respectively, for light, and circadian time 3 and 18, respectively, for serotonin. Quipazine, a nonspecific serotonin agonist, induced phase advance and phase delay at circadian time 3 and 18, respectively, like serotonin. (±)8-OH-DPAT, a specific 5-HT1A agonist, phase delayed by 2 h at the subjective night, but produced no significant phase shifts at the subjective day. When NAN-190, a specific 5-HT1A antagonist, was applied together with quipazine, it completely blocked the phase delay at circadian time 18, whereas it had no effect on the advance shifts induced by quipazine. The results suggest that the phase dependency of serotonin-induced phase shifts of the clock may be partly attributable to the daily change in receptor type. Accepted: 4 July 1999  相似文献   

5.
6.
MPer1 and mper2 are essential for normal resetting of the circadian clock   总被引:8,自引:0,他引:8  
Mammalian Per1 and Per2 genes are involved in the mechanism of the circadian clock and are inducible by light. A light pulse can evoke a change in the onset of wheel-running activity in mice by shifting the onset of activity to earlier times (phase advance) or later times (phase delays) thereby advancing or delaying the clock (clock resetting). To assess the role of mouse Per (mPer) genes in circadian clock resetting, mice carrying mutant mPer1 or mPer2 genes were tested for responses to a light pulse at ZT 14 and ZT 22, respectively. The authors found that mPer1 mutants did not advance and mPer2 mutants did not delay the clock. They conclude that the mammalian Per genes are not only light-responsive components of the circadian oscillator but also are involved in resetting of the circadian clock.  相似文献   

7.
The behavior of mammals is characterized by a 24-h cycle of rest and activity which is a fundamental adaption to the solar cycle of light and darkness. The pacemaker of this circadian clock is localized in the ventral part of the hypothalamus, the so-called suprachiasmatic nuclei (SCN), and is entrained by light signals mediated by the eye. The eye is directly connected via the retinohypothalamic tract (RHT) to the SCN. Light that reaches the retina elicits glutamate release at the synaptic terminals of the RHT and influences the neurons in the SCN in a manner that alters the behavioral state of the animal. A light pulse that reaches the retina at the beginning of the night elicits a delay of the clock phase, whereas a light pulse that reaches the retina at the end of the dark period leads to an advance of the clock phase. This advance or delay can be quantified by measuring the change in onset of wheel-running activity. Such measurements have, and continue to provide, a remarkably powerful assay of how light is detected and transduced to regulate circadian rhythms. The methods used for such measurements in mice are described in the following article.  相似文献   

8.
The discontinuous gas exchange cycle (DGC) was described in the German cockroach, Blattella germanica (L.) (Dictyoptera: Blattellidae) for the first time. Also, the effect of the DGC on water loss was investigated. The CO(2) emission pattern in both insecticide resistant and susceptible B. germanica varied with temperature. At 10, 15, and 20 degrees C the pattern was discontinuous. Cycle frequency increased at 25 and 30 degrees C, and at 35 degrees C the pattern became cyclic. In most DGCs, there was no clear distinction between the closed and flutter phases in both strains thus data for these phases were combined and analyzed as the interburst phase. The probability that B. germanica would breath discontinuously varied with temperature. Most cockroaches (62.8%) displayed DGCs at 10 degrees C, therefore measurement of metabolic rate and water loss was carried out at this temperature. Using repeated measures of analysis of variance, the interburst and burst V(.)(CO(2))(ml h(-1)) were not significantly different between the two strains. The variability in CO(2) emission during the interburst and burst phases over time was not significantly different from cycle to cycle or between strains. Overall metabolic rate during the entire recording was not significantly different between both strains. There was a significant difference in the duration of the interburst and burst phases between the strains. The susceptible strain had significantly longer interburst and burst phase durations during a complete DGC than the resistant strain. The interburst and burst phase durations were 5.01+/-0.19 and 6.21+/-0.13 min, respectively, for the resistant strain, whereas the durations were 7.16+/-0.37 and 6.73+/-0.17 min, respectively, for the susceptible strain. This resulted in a DGC of significantly longer duration (13.89+/-0.44 min) in the susceptible strain compared with the resistant strain (11.23+/-0.26 min). The duration of the interburst phase was significantly different from the open phase duration in the resistant strain such that during a single DGC lasting approximately 11.23 min, 43.5% consisted of the interburst phase while the burst phase made up 56.5% of the cycle. The cuticular permeability at 10 degrees C and 0% RH was 2.26 microg cm(-2) h(-1) mmHg(-1) for the resistant strain and 3.42 microg cm(-2) h(-1) mmHg(-1) for the susceptible strain. In both strains, cuticular transpiration accounted for approximately 95% of total water loss. The significantly longer duration of the interburst phase of the susceptible strain was not important in reducing water loss.  相似文献   

9.
Since the initial studies reporting that light can alter the phase position of the human circadian system, there has been increasing interest in the use of bright light as a tool for manipulating the phase position of the circadian pacemaker. Exposure protocols typically require subjects to receive 2-5 h of exposure over several circadian cycles. As a consequence, bright light treatment can involve a considerable time investment. However, recent studies indicate that a single pulse of bright light can produce significant phase shifts in the circadian pacemaker. If a single pulse of bright light can produce significant phase-shifting effects, multiple-pulse designs may be unnecessary. This study examined the phase-shifting effects of a single 4-h pulse of bright light (12,000 lux) in 14 male and one female subject aged between 19-45 years. With use of a “constant routine” to estimate circadian phase, a single 4-h pulse of light produced significant shifts in the phase of the core temperature rhythm. The timing of the exposure, relative to the core temperature rhythm, determined the degree and direction of the phase shift. Exposure immediately prior to habitual bedtime produced a mean phase delay in the core temperature of 2.39 h (SD = 1.37 h). In contrast, exposure immediately following habitual wake-up produced a mean phase advance of 1.49 h (SD = 2.06 h). In addition, the magnitude of the shift increased the closer the light pulse was to the individual's estimated endogenous core temperature minimum. There was, however, considerable interindividual variability in this relationship. Overall, these results confirm that a single pulse of bright light can produce significant phase shifts in the phase of the circadian pacemaker controlling core temperature. Key Words: Bright light—Circadian rhythm—Core body temperature—Sleep-wake disorders—Chronobiology.  相似文献   

10.
To survive, an animal must respond to the environmental stimuli (Zeitgebers) precisely coinciding with a 24‐h light/dark cycle. In human habitats, both natural and artificial Zeitgebers intermingle, entraining the circadian clocks of animals. Whether a circadian clock can respond to Zeitgebers effectively is closely related to the level of adaptive competitiveness. In the present review, the German cockroach (Blattella germanica L.) is compared with the double‐striped cockroach (Blattella bisignata Brunner von Wattenwyl) and the reasons why B. germanica is a more competitive dweller than its sibling species in human habitats are discussed. For example, sensitive adjustment of circadian clocks allows a male B. germanica to coordinate itself to the changes of daily light/dark cycles more efficiently; by contrast, it takes male B. bisignata more than 3 days to adjust. In a given dark phase, male B. germanica appear more to be resistant than B. bisignata to interference from light pulses. Ovary‐dependent masking factors enable female B. germanica to take advantage of the light phase to feed or mate, during which time a female B. bisignata typically remains inert. Furthermore, the gregariousness levels, fertilization and hatching rates, as well as mating behaviours of both cockroach species are compared. All of these comparisons suggest that B. germanica surpasses B. bisignata with respect to adapting to the artificial constraints within human habitats as a result of the adoption of an integrated strategy consisting of circadian and noncircadian approaches.  相似文献   

11.
Yellow wrasses (Halichoeres chrysus) show clear daily activity patterns. The fish hide in the substrate at (subjective) night, during the distinct rest phase. Initial entrainment in a 12h:12h light-dark (12:12 LD) cycle (mean period 24.02h, SD 0.27h, n = 16 was followed by a free run (mean period 24.42h, SD 1.33h) after transition into constant dim light conditions. Light pulses of a comparable intensity as used in the light part of the LD cycles did not result in significant phase shifts of the free-running rhythm in constant darkness. Application of much brighter 3h light pulses resulted in a phase-response curve (PRC) for a fish species, with pronounced phase advances during late subjective night. The PRCs differed from those mainly obtained in other vertebrate taxa by the absence of significant phase delays in the early subjective night. At that circadian phase, significant tonic effects of the light pulses caused a shortening of the circadian period length. Entrainment to skeleton photoperiods of 1:11 LD was observed in five of six wrasses exposed, also after a 3h phase advance of this LD cycle. Subsequently, a 1:11.25 LD cycle resulted in entrainment in four of the six fish. It is suggested that the expression of the circadian system in fish can be interpreted as a functional response to a weak natural zeitgeber, as present in the marine environment. This response allows photic entrainment as described here in the yellow wrasse. (Chronobiology International, 17(5), 613-622, 2000)  相似文献   

12.
Ramelteon, an MT(1)/MT(2) melatonin receptor agonist, is used for the treatment of sleep-onset insomnia and circadian sleep disorders. Ramelteon phase shifts circadian rhythms in rodents and humans when given at the end of the subjective day; however, its efficacy at other circadian times is not known. Here, the authors determined in C3H/HeN mice the maximal circadian sensitivity for ramelteon in vivo on the onset of circadian running-wheel activity rhythms, and in vitro on the peak of circadian rhythm of neuronal firing in suprachiasmatic nucleus (SCN) brain slices. The phase response curve (PRC) for ramelteon (90?μg/mouse, subcutaneous [sc]) on circadian wheel-activity rhythms shows maximal sensitivity during the late mid to end of the subjective day, between CT8 and CT12 (phase advance), and late subjective night and early subjective day, between CT20 and CT2 (phase delay), using a 3-day-pulse treatment regimen in C3H/HeN mice. The PRC for ramelteon resembles that for melatonin in C3H/HeN mice, showing the same magnitude of maximal shifts at CT10 and CT2, except that the range of sensitivity for ramelteon (CT8-CT12) during the subjective day is broader. Furthermore, in SCN brain slices in vitro, ramelteon (10 pM) administered at CT10 phase advances (5.6?±?0.29?h, n?=?3) and at CT2 phase delays (-3.2?±?0.12?h, n?=?6) the peak of circadian rhythm of neuronal firing, with the shifts being significantly larger than those induced by melatonin (10 pM) at the same circadian times (CT10: 2.7?±?0.15?h, n?=?4, p?相似文献   

13.
The suprachiasmatic nuclei of the hypothalamus contain the major circadian pacemaker in mammals, driving circadian rhythms in behavioral and physiological functions. This circadian pacemaker's responsiveness to light allows synchronization to the light-dark cycle. Phase shifting by light often involves several transient cycles in which the behavioral activity rhythm gradually shifts to its steady-state position. In this article, the authors investigate in Syrian hamsters whether a phase-advancing light pulse results in immediate shifts of the PRC at the next circadian cycle. In a first series of experiments, the authors aimed a light pulse at CT 19 to induce a phase advance. It appeared that the steady-state phase advances were highly correlated with activity onset in the first and second transient cycle. This enabled them to make a reliable estimate of the steady-state phase shift induced by a phase-advancing light pulse on the basis of activity onset in the first transient cycle. In the next series of experiments, they presented a light pulse at CT 19, which was followed by a second light pulse aimed at the delay zone of the PRC on the next circadian cycle. The immediate and steady-state phase delays induced by the second light pulse were compared with data from a third experiment in which animals received a phase-delaying light pulse only. The authors observed that the waveform of the phase-delay part of the PRC (CT 12-16) obtained in Experiment 2 was virtually identical to the phase-delay part of the PRC for a single light pulse (obtained in Experiment 3). This finding allowed for a quantitative assessment of the data. The analysis indicates that the delay part of the PRC-between CT 12 and CT 16-is rapidly reset following a light pulse at CT 19. These findings complement earlier findings in the hamster showing that after a light pulse at CT 19, the phase-advancing part of the PRC is immediately shifted. Together, the data indicate that the basis for phase advancing involves rapid resetting of both advance and delay components of the PRC.  相似文献   

14.
Although light is considered the primary entrainer of circadian rhythms in humans, nonphotic stimuli, including exercise and melatonin also phase shift the biological clock. Furthermore, in birds and nonhuman mammals, auditory stimuli are effective zeitgebers. This study investigated whether a nonphotic auditory stimulus phase shifts human circadian rhythms. Ten subjects (5 men and 5 women, ages 18-72, mean age +/- SD, 44.7 +/- 21.4 yr) completed two 4-day laboratory sessions in constant dim light (<20 lux). They received two consecutive presentations of either a 2-h auditory or control stimulus from 0100 to 0300 on the second and third nights (presentation order of the stimulus and control was counterbalanced). Core body temperature (CBT) was collected and stored in 2-min bins throughout the study and salivary melatonin was obtained every 30 min from 1900 to 2330 on the baseline and poststimulus/postcontrol nights. Circadian phase of dim light melatonin onset (DLMO) and of CBT minimum, before and after auditory or control presentation was assessed. The auditory stimulus produced significantly larger phase delays of the circadian melatonin (mean +/- SD, -0.89 +/- 0.40 h vs. -0.27 +/- 0.16 h) and CBT (-1.16 +/- 0.69 h vs. -0.44 +/- 0.27 h) rhythms than the control. Phase changes for the two circadian rhythms also positively correlated, indicating direct effects on the biological clock. In addition, the auditory stimulus significantly decreased fatigue compared with the control. This study is the first demonstration of an auditory stimulus phase-shifting circadian rhythms in humans, with shifts similar in size and direction to those of other nonphotic stimuli presented during the early subjective night. This novel stimulus may be a useful countermeasure to facilitate circadian adaptation after transmeridian travel or shift work.  相似文献   

15.
Male adult German cockroaches, Blattella germanica (L.), expressed robust locomotor circadian rhythmicity under 28 degrees C and constant darkness (DD) conditions. By surgically severing the connections between the optic lobes and midbrain and their subsequent regeneration, we demonstrated that the locomotor circadian pacemaker was located in the optic lobes and that it controlled the locomotor circadian rhythm through neural pathways. From the results that unilaterally optic tract severed males still showed locomotor circadian rhythmicity (93.1%, n=29) without significantly changing the circadian period (tau) or level of motor activity, we concluded that the right and left optic lobes each contain a circadian pacemaker competent to drive the locomotor circadian rhythm. These two pacemakers were strongly coupled since only one rhythm was expressed when the pacemakers were independently exposed to opposite lighting conditions (DD or LL) at the same time. However, an unequal contribution of each pacemaker in determining the overt circadian period was found under constant dim light (10 lux) conditions, revealing a major-minor coupling relationship between these two pacemakers, so that the unilaterally blinded male expressed either a LL-rhythm with a circadian period of 24.27+/-0.21 h (41.7%, n=24) or a DD-rhythm with a circadian period of 23.43+/-0.19 h (58.3%, n=24). However, higher intensity of photic information (200-300 lux) could overpower this relationship and cause the minor pacemaker to lead the rhythm.  相似文献   

16.
Mammalian circadian rhythms are generated by a hypothalamic suprachiasmatic nuclei (SCN) clock. Light pulses synchronize body rhythms by inducing phase delays during the early night and phase advances during the late night. Phosphorylation events are known to be involved in circadian phase shifting, both for delays and advances. Pharmacological inhibition of the cGMP-dependent kinase (cGK) or Ca2+/calmodulin-dependent kinase (CaMK), or of neuronal nitric oxide synthase (nNOS) blocks the circadian responses to light in vivo. Light pulses administered during the subjective night, but not during the day, induce rapid phosphorylation of both p-CAMKII and p-nNOS (specifically phosphorylated by CaMKII). CaMKII inhibitors block light-induced nNOS activity and phosphorylation, suggesting a direct pathway between both enzymes. Furthermore, SCN cGMP exhibits diurnal and circadian rhythms with maximal values during the day or subjective day. This variation of cGMP levels appears to be related to temporal changes in phosphodiesterase (PDE) activity and not to guanylyl cyclase (GC) activity. Light pulses increase SCN cGMP levels at circadian time (CT) 18 (when light causes phase advances of rhythms) but not at CT 14 (the time for light-induced phase delays). cGK II is expressed in the hamster SCN and also exhibits circadian changes in its levels, peaking during the day. Light pulses increase cGK activity at CT 18 but not at CT 14. In addition, cGK and GC inhibition by KT-5823 and ODQ significantly attenuated light-induced phase shifts at CT 18. This inhibition did not change c-Fos expression SCN but affected the expression of the clock gene per in the SCN. These results suggest a signal transduction pathway responsible for light-induced phase advances of the circadian clock which could be summarized as follows: Glu-Ca2+-CaMKII-nNOS-GC-cGMP-cGK-->-->clock genes. This pathway offers a signaling window that allows peering into the circadian clock machinery in order to decipher its temporal cogs and wheels.  相似文献   

17.
Since the initial studies reporting that light can alter the phase position of the human circadian system, there has been increasing interest in the use of bright light as a tool for manipulating the phase position of the circadian pacemaker. Exposure protocols typically require subjects to receive 2–5 h of exposure over several circadian cycles. As a consequence, bright light treatment can involve a considerable time investment. However, recent studies indicate that a single pulse of bright light can produce significant phase shifts in the circadian pacemaker. If a single pulse of bright light can produce significant phase-shifting effects, multiple-pulse designs may be unnecessary. This study examined the phase-shifting effects of a single 4-h pulse of bright light (12,000 lux) in 14 male and one female subject aged between 19–45 years. With use of a “constant routine” to estimate circadian phase, a single 4-h pulse of light produced significant shifts in the phase of the core temperature rhythm. The timing of the exposure, relative to the core temperature rhythm, determined the degree and direction of the phase shift. Exposure immediately prior to habitual bedtime produced a mean phase delay in the core temperature of 2.39 h (SD = 1.37 h). In contrast, exposure immediately following habitual wake-up produced a mean phase advance of 1.49 h (SD = 2.06 h). In addition, the magnitude of the shift increased the closer the light pulse was to the individual's estimated endogenous core temperature minimum. There was, however, considerable interindividual variability in this relationship. Overall, these results confirm that a single pulse of bright light can produce significant phase shifts in the phase of the circadian pacemaker controlling core temperature.  相似文献   

18.
Melatonin is known to shift the phase of the locomotor activity rhythm in the field mouse Mus booduga in accordance with a type-I phase response curve (PRC), with phase delays during the subjective day and phase advances during late subjective night and the early subjective day. At CT4 (circadian time 4; i.e. 16 hr. after activity onset) and CT22 of the circadian cycle, a single dose of melatonin (1 mg/kg) is known to evoke maximum delay and maximum advance phase-shifts, respectively. We investigated the dose-dependent responses of the circadian pacemaker of these mice to a single dose of melatonin at the times for maximum delay and maximum advance. The circadian pacemaker responsible for the locomotor activity rhythm in these mice responded to various doses of melatonin in a dose-dependent manner with the magnitude of phase shifts increasing with dose.  相似文献   

19.
Melatonin is known to shift the phase of the locomotor activity rhythm in the field mouse Mus booduga in accordance with a type-I phase response curve (PRC), with phase delays during the subjective day and phase advances during late subjective night and the early subjective day. At CT4 (circadian time 4; i.e. 16 hr. after activity onset) and CT22 of the circadian cycle, a single dose of melatonin (1 mg/kg) is known to evoke maximum delay and maximum advance phase-shifts, respectively. We investigated the dose-dependent responses of the circadian pacemaker of these mice to a single dose of melatonin at the times for maximum delay and maximum advance. The circadian pacemaker responsible for the locomotor activity rhythm in these mice responded to various doses of melatonin in a dose-dependent manner with the magnitude of phase shifts increasing with dose.  相似文献   

20.
The effect of altitude on four basic properties of the pacemaker controlling the circadian rhythm of oviposition in two strains of Drosophila ananassae was determined. The high altitude (HA) strain from Badrinath (5123 m above sea level) had a low amplitude peak in the forenoon while the low altitude (LA) strain from Firozpur (179 m a.s.l.) had a high amplitude peak after the lights-off of LD 12:12 cycles. Free running periods in continuous darkness were about 22.6 and 27.4 h in the HA and LA strains, respectively. The light pulse phase response curve (PRC) for the HA strain showed a low amplitude and a dead zone of 8h; the ratio for the advance to delay region (A/D) was less than 1, while the PRC for the LA strain had a high amplitude, which was devoid of a dead zone and showed a ratio of A/D > 1. The magnitude of the delay phase shifts at CT 18 evoked by light pulses of 1 h duration, but varying light intensity was significantly different in the HA and LA strain, which suggests that the photic sensitivity of the clock photoreceptors mediating the phase shifts had been affected by the altitude.  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