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.
Trace eyeblink conditioning (with a trace interval ≥500 msec) depends on the integrity of the hippocampus and requires that participants develop awareness of the stimulus contingencies (i.e., awareness that the conditioned stimulus [CS] predicts the unconditioned stimulus [US]). Previous investigations of the relationship between trace eyeblink conditioning and awareness of the stimulus contingencies have manipulated awareness or have assessed awareness at fixed intervals during and after the conditioning session. In this study, we tracked the development of knowledge about the stimulus contingencies trial by trial by asking participants to try to predict either the onset of the US or the onset of their eyeblinks during differential trace eyeblink conditioning. Asking participants to predict their eyeblinks inhibited both the acquisition of awareness and eyeblink conditioning. In contrast, asking participants to predict the onset of the US promoted awareness and facilitated conditioning. Acquisition of knowledge about the stimulus contingencies and acquisition of differential trace eyeblink conditioning developed approximately in parallel (i.e., concurrently).  相似文献   

2.
We examined the importance of awareness for eyeblink conditioning by directly comparing singlecue delay eyeblink conditioning and single-cue trace eyeblink conditioning. During single-cue delay conditioning, participants who became aware of the stimulus contingencies early in the conditioning session conditioned no better than those who became aware later in the session or did not become aware. Thus, the level of awareness was unrelated to the overall level of conditioning across the session. In contrast, awareness of the stimulus contingencies early in the session predicted the success of single-cue trace conditioning. These data, together with earlier findings, show that awareness is irrelevant to single-cue delay eyeblink conditioning but is critical for single-cue trace eyeblink conditioning. The findings from the present study are related to previous findings for differential (CS+ and CS) eyeblink conditioning and awareness.  相似文献   

3.
P. F. Lovibond and D. R. Shanks (2002) suggested that all forms of classical conditioning depend on awareness of the stimulus contingencies. This article considers the available data for eyeblink classical conditioning, including data from 2 studies (R. E. Clark, J. R. Manns, & L. R. Squire, 2001; J. R. Manns, R. E. Clark, & L. R. Squire, 2001) that were completed too recently to have been considered in their review. In addition, in response to questions raised by P. F. Lovibond and D. R. Shanks, 2 new analyses of data are presented from studies published previously. The available data from humans and experimental animals provide strong evidence that delay eyeblink classical conditioning (but not trace eyeblink classical conditioning) can be acquired and retained independently of the forebrain and independently of awareness. This conclusion applies to standard conditioning paradigms; for example, to single-cue delay conditioning when a tone is used as the conditioned stimulus (CS) and to differential delay conditioning when the positive and negative conditioned stimuli (CS+ and CS-) are a tone and white noise.  相似文献   

4.
In rodents stress impairs delay as well as trace eyelid conditioning in females, but enhances it in males. The present study tested the effects of acute psychosocial stress exposure on classical delay eyeblink conditioning in healthy men and women. In a between subject design, participants were exposed to psychosocial stress using the Trier Social Stress Test (TSST) or a control condition which was followed by a delay eyeblink classical conditioning procedure. Stress exposure led to a significant increase in salivary cortisol and impaired acquisition of conditioned eyeblink responses (CRs). This was evident by a later first CR and an overall lower CR rate of the stress group. The stress-induced acquisition impairment was observed in both women and men. Subjects failing to show a stress-induced cortisol increase (cortisol non-responder) were not impaired in acquisition. Our findings indicate that acute stress, possibly via activation of the hypothalamus–pituitary–adrenal (HPA) axis, reduces the ability to acquire a simple conditioned motor response in humans.  相似文献   

5.
Delay eyeblink conditioning is established by paired presentations of a conditioned stimulus (CS) such as a tone or light, and an unconditioned stimulus (US) that elicits the blink reflex. Conditioned stimulus information is projected from the basilar pontine nuclei to the cerebellar interpositus nucleus and cortex. The cerebellar cortex, particularly the molecular layer, contains a high density of cannabinoid receptors (CB1R). The CB1Rs are located on the axon terminals of parallel fibers, stellate cells, and basket cells where they inhibit neurotransmitter release. The present study examined the effects of a CB1R agonist WIN55,212-2 and antagonist SR141716A on the acquisition of delay eyeblink conditioning in rats. Rats were given subcutaneous administration of 1, 2, or 3 mg/kg of WIN55,212-2 or 1, 3, or 5 mg/kg of SR141716A before each day of acquisition training (10 sessions). Dose-dependent impairments in acquisition were found for WIN55,212-2 and SR141716A, with no effects on spontaneous or nonassociative blinking. However, the magnitude of impairment was greater for WIN55,212-2 than SR141716A. Dose-dependent impairments in conditioned blink response (CR) amplitude and timing were found with WIN55,212-2 but not with SR141716A. The findings support the hypothesis that CB1Rs in the cerebellar cortex play an important role in plasticity mechanisms underlying eyeblink conditioning.  相似文献   

6.
7.
Pavlovian eyeblink conditioning has been used extensively as a model system for examining the neural mechanisms underlying associative learning. Delay eyeblink conditioning depends on the intermediate cerebellum ipsilateral to the conditioned eye. Evidence favors a two-site plasticity model within the cerebellum with long-term depression of parallel fiber synapses on Purkinje cells and long-term potentiation of mossy fiber synapses on neurons in the anterior interpositus nucleus. Conditioned stimulus and unconditioned stimulus inputs arise from the pontine nuclei and inferior olive, respectively, converging in the cerebellar cortex and deep nuclei. Projections from subcortical sensory nuclei to the pontine nuclei that are necessary for eyeblink conditioning are beginning to be identified, and recent studies indicate that there are dynamic interactions between sensory thalamic nuclei and the cerebellum during eyeblink conditioning. Cerebellar output is projected to the magnocellular red nucleus and then to the motor nuclei that generate the blink response(s). Tremendous progress has been made toward determining the neural mechanisms of delay eyeblink conditioning but there are still significant gaps in our understanding of the necessary neural circuitry and plasticity mechanisms underlying cerebellar learning.  相似文献   

8.
Big-K+ conductance (BK)-channel mediated fast afterhyperpolarizations (AHPs) following action potentials are reduced after eyeblink conditioning. Blocking BK channels with paxilline increases evoked firing frequency in vitro and spontaneous pyramidal activity in vivo. To examine how increased excitability after BK-channel blockade affects learning, rats received bilateral infusions of paxilline, saline, or nothing into hippocampal CA1 prior to trace eyeblink conditioning. The drug group was slower to acquire the task, but learning was not completely impaired. This suggests that nonspecific increases in excitability and baseline neuronal firing rates caused by in vivo blockade of the BK channel may disrupt correct processing of inputs, thereby impairing hippocampus-dependent learning.Learning and increased neuronal intrinsic excitability are strongly correlated, although a causal relationship has not yet been definitively established (Disterhoft and Oh 2006). One of the mechanisms of increased excitability is through reduction of potassium currents, which cause afterhyperpolarizations (AHP). Afterhyperpolarizations in pyramidal cells can be divided into three categories based upon their timing and duration. The fast AHP lasts only 2–5 ms, follows the depolarizing phase of individual action potentials, and is mediated largely by the big-K+ conductance (BK) channel. The post-burst AHP has a medium (50–100 ms) and a slow (1–2 s) component, and follows trains or bursts of action potentials. The medium AHP is carried by apamin sensitive small-K+ conductance (SK) channels, but the channel(s), which carry the slow AHP, are still unknown (Storm 1987; Disterhoft and Oh 2006).Learning-related reductions in the post-burst AHP are well documented (for review see Disterhoft and Oh 2006). Additionally, pharmacological modulators of the post-burst AHP alter learning in an expected manner—compounds that reduce the AHP improve learning (galantamine [Simon et al. 2004] and nimodopine [Deyo et al. 1989]). Learning-related reductions of the fast AHP are also seen in prefrontal cortex pyramidal neurons after extinction of fear conditioning (Santini et al. 2008) and in CA1 hippocampal pyramidal neurons after learning trace eyeblink conditioning (tEBC) (Matthews et al. 2008). In vitro whole-cell recordings show that blocking the BK channel with either paxilline or iberiotoxin increases the firing rate to a step current injection (Nelson et al. 2003). Likewise, injection of paxilline into the hippocampus increases the in vivo spontaneous firing frequency of hippocampal CA1 neurons of awake freely moving rats up to 2.5-fold over saline injections (Matthews et al. 2008), indicating that the BK-mediated fast AHP plays an important role in intrinsic excitability. The current study was undertaken to determine whether pharmacologically reducing the fast AHP during training would improve trace eyeblink conditioning.Experimental subjects were 3- to 4-mo old Fisher 344 X Brown Norway F1-hybrid rats. Animals were housed in pairs, in a climate-controlled facility with a 12:12 light–dark cycle, and ad libitum access to food and water. Procedures were in accordance with the guidelines of the Northwestern University Animal Care and Use Committee and conformed to NIH standards (NIH Publication No. 80-23). All efforts were made to minimize the number of animals utilized and their discomfort. Thirty-seven rats were originally included in the study, however 13 were excluded from the experiment due to incorrect cannulae placement, faulty EMG signal, or unrelated health issues. The final groups included in the study were nine drug animals, seven vehicle animals, eight sham animals, and six non-cannulated animals.Guide cannulae (26 gauge stainless steel) were bilaterally implanted at −3.6 mm AP, ±2.0 mm ML. The guide cannulae were lowered slowly (0.5 mm/5 min) to a depth of −1.9 mm subdura. The cannulae were cemented in place with dental acrylic. An electronic connector strip with pins for ground and two EMG wires was fitted between the two guide cannulae and grounded to the skull screws. The EMG wires were implanted under the right eyelid and the entire apparatus was cemented in place (Weiss et al. 1999). Rats were given Buprenex (0.05 mg/kg) post-surgery to alleviate any discomfort.Cannulae placements were verified histologically after training was completed. Animals were given a lethal dose of barbiturate (0.15 mL i.p.) then transcardially perfused with 0.9% saline followed by 10% formalin. After perfusion, the brains were carefully immersed in 10% formalin for a minimum of 24 h. Eighty micron coronal sections were made with a freezing cryostat, and every second section was kept and mounted on gelatin coated slides. Sections were stained with cresyl violet to reveal cell death or excessive damage surrounding the injection cannulae. Animals with incorrectly placed cannulae or excessive tissue damage were not included in the study. A diagram of the most ventral extent of each cannula is shown in Figure 1A. Paxilline, a BK-channel blocker, is a peptide and there is little known about the motility of this molecule in the brain. From previous in vivo recording work (Matthews et al. 2008), it is known that the BK blocker is active and able to diffuse at least a radius of 1.7 mm from the tip of the cannula. To approximate the spread of paxilline in the present study, 1.0% ibotenic acid was injected into two animals at the completion of training in a manner that exactly mimicked the paxilline injections in volume and rate. Five days after excitotoxic lesion, animals were sacrificed and their brains processed. Figure 1B shows the maximum and minimum extent of cell loss due to ibotenic acid lesion.Open in a separate windowFigure 1.Cannula placement and drug spread. (A) Cannulae were bilaterally implanted to terminate directly above the CA1 layer of the hippocampus. Placement was verified after the behavioral experiments. Gray dots indicate the location of the tip of each cannula for animals included in the study. (B) The spread of the drug was approximated by injecting 1.0% ibotenic acid into the left hemisphere of two animals at the end of training. The injection volume and rate were the same as those used for the study. The right hemisphere served as a within-animal control for the action of the ibotenic acid. The maximal (light gray) and minimal (dark gray) spread are shown. Measurements are relative to bregma. (Adapted from Paxinos and Watson [1998] and reprinted with permission from Elsevier ©1998.)Freely moving animals were injected with vehicle (1% DMSO in saline), drug (1 μM paxilline in saline), or nothing 20–30 min before the start of training, including the first habituation session. The trainer was blind to the identity of the injected substance. Infusions were performed using two 2 μL Hamilton syringes connected by lengths of flexible, oil filled tubing to 33 gauge infusion needles, which extended 0.5 mm beyond the end of the guide cannulae. One microliter of sterile solution was infused into each hemisphere at a constant rate of 0.2 μL/min using a Stoelting double-barrel infusion pump. The injection needles were left in place for 1 min following the injection to allow diffusion away from the tip of the needle.Trace eyeblink conditioning is an associative learning task in which a neutral conditioned stimulus (CS) is paired with an unconditioned stimulus (US) that elicits a reflexive eyelid closure. The insertion of a stimulus-free “trace” interval between the CS and the US makes this task strongly dependent on the hippocampus (Solomon et al. 1986; Moyer Jr. et al. 1990). After repeated pairings of the CS and the US, if an association has been learned, the subject will begin to blink during the trace period in anticipation of the US. Trace eyeblink conditioning procedures as described by Weiss et al. (1999) were followed. Training sessions were conducted in a sound-attenuating chamber and controlled with a custom-designed LabVIEW (National Instruments) program; eyelid EMG data were integrated online during training. Animals were connected to the recording computer via the implanted connector strip; a short tether served the dual purpose of allowing EMG activity to be monitored and positioning an air puff delivery tube in front of the eye, while the rat was freely moving. On the first day, the subjects received a session of stimulus-free habituation to the training chamber lasting as long as a conditioning session. The subsequent 5 d were training sessions. Conditioned animals received 30 trials per session (30 s average ITI) for a total of 150 CS–US pairings, consisting of a tone stimulus (CS, 80 dB, 250 ms) paired with a corneal air puff (US, 3–5 psi, 100 ms) with a 250 ms stimulus-free trace interval interposed.The primary measure of tEBC learning used in this study was a correctly timed eyelid closure, i.e., an eyelid closure that begins during the trace interval and continues until the air puff. Eyelid activity was measured with an implanted EMG electrode. The division of eyelid responses into “adaptive” and “nonadaptive” categories has been used in other studies (Garcia et al. 1999). For this reason, we analyzed eyelid closure during the entire tone/trace period and during only the last 200 ms preceding the air puff. Responses given during the 200 ms preceding the US are termed adaptive conditioned responses (CR). Figure 2 shows the timing of the tone, trace, and air puff; the timeline for each type of response; and an integrated EMG. Any trials in which the baseline activity in the 500 ms preceding CS presentation exceeded two standard deviations were discarded. Eye closure was defined as greater than four standard deviations above baseline. Averages for all relevant measures for each session were compared between groups for training sessions 1–5 using repeated-measures ANOVA. Learning across training sessions within each group was assessed with a planned comparison ANOVA using StatVIEW software. The stimulus-free habituation session was excluded from all between-group analyses. The percentage of aCRs during habituation is shown in Figure 3 to provide the baseline level of spontaneous eyelid closures.Open in a separate windowFigure 2.Timing of conditioned responses (CRs) in relation to the tone, trace, and air puff. The integrated EMG shows a robust adaptive eyelid closure in anticipation of the air puff. The tone lasted 250 ms and was followed by a 250 ms trace period. The air puff was 100 ms long. Below the EMG trace is a time line showing the period for an adaptive CR and an unconditioned response (UR).Open in a separate windowFigure 3.BK block slows learning of trace eyeblink conditioning. Vehicle, drug, and sham animals were injected and immediately trained on the trace eyeblink task over 6 d, one session each day. (Non-cannulated animals were trained in two sessions per day for 3 d.) The first session for all groups was stimuli-free habituation. Each session consisted of 30 pairings of a CS tone with a US air puff. Injection of the BK blocker paxilline resulted in slower acquisition of the task (*P < 0.05), although animals in all groups were eventually able to learn the task (#P < 0.001).The concentration of paxilline was selected based on the dose that achieved a maximal increase in spontaneous firing in vivo (Matthews et al. 2008). Two groups were designed to control for pressure effects of the injection or the stress and tissue damage of the cannulation surgery. These were a vehicle group (1% DMSO in saline) and a sham group (empty needles). In addition, the learning behavior of these cannulated animals was compared to a separate group of animals that had only the EBC head apparatus implanted (non-cannulated).Rats that received an infusion of paxilline (1 μM) immediately preceding training were significantly slower to acquire the task as measured by the percent of adaptive CRs exhibited across all training sessions (F (3,26) = 3.155, P = 0.042, repeated-measures ANOVA, Fisher''s PLSD, P < 0.02) (Fig. 3). The percentage of responses during the entire tone/trace period showed a trend toward reduced learning in the drug group (F (3,26) = 2.729, P = 0.064, repeated-measures ANOVA).Other parameters of the eyelid closure were examined. There were no significant differences in the peak, onset, duration, or area of the adaptive CR. There was also no difference between groups in the onset of the air-puff elicited eyelid closure (unconditioned response [UR]), suggesting that the drug did not cause decreased sensitivity to the US. Finally, the baseline eyelid EMG activity of all four groups during the stimulus-free habituation session was not significantly different, suggesting that the drug did not suppress or enhance spontaneous eyelid activity.All animals in the study showed improved performance across the training sessions (F (4,104) = 21.810, P < 0.0001, repeated-measures ANOVA). Further analyses revealed that animals in the drug group also showed continuous learning over the training days (Session 1: 16.1% ± 4.9%, Session 5: 56.8% ± 8.2%; F (4,32) = 6.961; P = 0.0004), and eventually the drug group reached a percentage of adaptive CRs statistically comparable to the controls (F (3,26) = 2.377, P = 0.093 for Session 5, ANOVA).There is a precedent for anticipating enhanced learning after in vivo pharmacological manipulations of intrinsic excitability (Disterhoft and Oh 2006). The long-lasting post-burst AHP is reduced in hippocampal cells from animals that have learned a hippocampus-dependent task (Moyer Jr. et al. 1996; Oh et al. 2003), and increased in aging animals that have difficulty learning. Pharmacologically reducing the post-burst AHP in vivo with calcium channel blockers (nimodipine) (Deyo et al. 1989), acetylcholine agonists (CI-1017) (Weiss et al. 2000), or cholinesterase inhibitors (galantamine and metrifonate) (Kronforst-Collins et al. 1997; Weible et al. 2004) leads to improved learning in aging animals. Thus, the finding that blocking the BK channel results in slowed learning is somewhat surprising, given the previous report of a reduction in the BK-mediated fast AHP after learning tEBC (Matthews et al. 2008). Several explanations may account for why pharmacologically reducing the fast AHP in vivo impaired rather than improved learning.First, the reduction of the fast AHP seen with channel blockers in in vitro experiments is of greater magnitude than the reductions in the fast AHP seen after learning; additionally, the fast AHP in cells from trained animals can be further reduced in vitro with paxilline or iberiotoxin (Matthews et al. 2008). It could be that there is an important difference between reducing the BK-mediated current, as is seen after learning, and completely blocking it using a drug. Reducing the fast AHP increases intrinsic excitability, and completely blocking the BK channel increases excitability even further. However, it is important to emphasize recent research showing how excessive excitability can be detrimental to learning. An early indicator of mild cognitive impairment detected with fMRI is hyperactivity in the hippocampus and medial temporal lobe (Miller et al. 2008). At the cellular level, saturating in vivo hippocampal LTP results in impaired spatial learning due to increased epileptiform activity, rather than saturated synaptic plasticity (McNamara et al. 1993). Finally, work with a knockout model of the β-4 subunit of the BK channel shows that this calcium- and voltage-dependent channel helps to regulate hyperexcitability and reduce the occurrence of temporal lobe seizures (Brenner et al. 2005). Although intrahippocampal paxilline infusion did not cause epileptiform activity in in vivo recordings, or observable seizure behavior (Juhng et al. 1999), it is possible that the drug caused pathological increases in excitability that impeded learning.Second, unregulated or meaningless increases in baseline neural activity in the hippocampus increase background noise, effectively decreasing the signal-to-noise ratio for the whole network, making it more difficult to distinguish important, information-bearing activity from background noise. In delay conditioning (where there is no trace interval between the CS and US), ablation of the hippocampus does not disrupt eyeblink conditioning (Solomon et al. 1986; Hesslow 1994), however, increasing (Salafia et al. 1979) or suppressing (Solomon et al. 1983) hippocampal activity has a strong retarding effect on learning this task. Disrupting synaptic transmission in a subregion of the hippocampus also impairs spatial learning (Niewoehner et al. 2007), further illustrating how a disturbance of information processing at a single junction of the trisynaptic circuit can impair learning. BK channels are present in presynaptic terminals, as well as at the soma, where they participate in controlling transmitter release. Blocking BK channels decreases failures and increases the amplitude of EPSCs at CA3–CA3 synapses (Raffaelli et al. 2004). The infusion of paxilline may have also increased the efficacy and frequency of transmitter release at the CA3–CA1 synapse in the present experiment. It may be that by selecting a dose of paxilline that caused a maximum increase in in vivo spontaneous activity, we overdosed the hippocampus to levels of excitability that interfered with stimulus processing, thereby impairing and slowing learning.Finally, the fast AHP is largely mediated by the BK channel, but other potassium currents also play a role in action potential repolarization. The A-type potassium current in particular has been implicated in learning-related excitability changes and is active during an action potential (Giese et al. 1998). Changing action potential repolarization dynamics also alters the calcium influx into cells (Zhou et al. 2005), which can have far reaching effects on other calcium-dependent processes, such as gene regulation, synaptic plasticity, or protein expression. The learning impairment seen in this study might be due to secondary effects on other potassium currents or calcium-dependent processes.This study indicates that normal activity of BK channels is required for acquisition of the tEBC task. The channel may act to maintain neurons within a narrow window of excitability, keeping neurons within an operating range of “optimal excitability.” BK channel activity is strongly modulated by kinase-phosphotase activity (Reinhart et al. 1991; Loane et al. 2006), and reduction of BK-mediated current through modulators of these molecules may have a more beneficial impact on learning. Since blocking of the BK channel with paxilline impedes learning, this drug may not present a useful tool for pharmacological learning-enhancement manipulations.  相似文献   

9.
The auditory conditioned stimulus (CS) pathway that is necessary for delay eyeblink conditioning was investigated using reversible inactivation of the medial auditory thalamic nuclei (MATN) consisting of the medial division of the medial geniculate (MGm), suprageniculate (SG), and posterior intralaminar nucleus (PIN). Rats were given saline or muscimol infusions into the MATN contralateral to the trained eye before each of four conditioning sessions with an auditory CS. Rats were then given four additional sessions without infusions to assess savings from the initial training. All rats were then given a retention test with a muscimol infusion followed by a recovery session. Muscimol infusions through cannula placements within 0.5 mm of the MGm prevented acquisition of eyeblink conditioned responses (CRs) and also blocked CR retention. Cannula placements more than 0.5 mm from the MATN did not completely block CR acquisition and had a partial effect on CR retention. The primary and secondary effects of MATN inactivation were examined with 2-deoxy-glucose (2-DG) autoradiography. Differences in 2-DG uptake in the auditory thalamus were consistent with the cannula placements and behavioral results. Differences in 2-DG uptake were found between groups in the ipsilateral auditory cortex, basilar pontine nuclei, and inferior colliculus. Results from this experiment indicate that the MATN contralateral to the trained eye and its projection to the pontine nuclei are necessary for acquisition and retention of eyeblink CRs to an auditory CS.  相似文献   

10.
Cholinergic systems are critical to the neural mechanisms mediating learning. Reduced nicotinic cholinergic receptor (nAChR) binding is a hallmark of normal aging. These reductions are markedly more severe in some dementias, such as Alzheimer's disease. Pharmacological central nervous system therapies are a means to ameliorate the cognitive deficits associated with normal aging and aging-related dementias. Trace eyeblink conditioning (EBC), a hippocampus- and forebrain-dependent learning paradigm, is impaired in both aged rabbits and aged humans, attributable in part to cholinergic dysfunction. In the present study, we examined the effects of galantamine (3 mg/kg), a cholinesterase inhibitor and nAChR allosteric potentiating ligand, on the acquisition of trace EBC in aged (30–33 months) and young (2–3 months) female rabbits. Trace EBC involves the association of a conditioned stimulus (CS) with an unconditioned stimulus (US), separated by a stimulus-free trace interval. Repeated CS–US pairings results in the development of the conditioned eyeblink response (CR) prior to US onset. Aged rabbits receiving daily injections of galantamine (Aged/Gal) exhibited significant improvements compared with age-matched controls in trials to eight CRs in 10 trial block criterion (P = 0.0402) as well as performance across 20 d of training [F(1,21) = 5.114, P = 0.0345]. Mean onset and peak latency of CRs exhibited by Aged/Gal rabbits also differed significantly [F(1,21) = 6.120/6.582, P = 0.0220/0.0180, respectively] compared with age-matched controls, resembling more closely CR timing of young drug and control rabbits. Galantamine did not improve acquisition rates in young rabbits compared with age-matched controls. These data indicate that by enhancing nicotinic and muscarinic transmission, galantamine is effective in offsetting the learning deficits associated with decreased cholinergic transmission in the aging brain.  相似文献   

11.
The essential neural circuitry for delay eyeblink conditioning has been largely identified, whereas much of the neural circuitry for trace conditioning has not been identified. The major difference between delay and trace conditioning is a time gap between the presentation of the conditioned stimulus (CS) and the unconditioned stimulus (US) during trace conditioning. It is this time gap or trace interval which accounts for an additional memory component in trace conditioning. Additional neural structures are also necessary for trace conditioning, including hippocampus and prefrontal cortex. This addition of forebrain structures necessary for trace but not delay conditioning suggests other brain areas become involved when a memory gap is added to the conditioning parameters. A metabolic marker of energy use, radioactively labeled glucose analog, was used to compare differences in glucose analog uptake between delay, trace, and unpaired experimental groups in order to identify new areas of involvement within the cerebellum. Known structures such as the interpositus nucleus and lobule HVI showed increased activation for both delay and trace conditioning compared to unpaired conditioning. However, there was a differential amount of activation between anterior and posterior portions of the interpositus nucleus between delay and trace, respectively. Cerebellar cortical areas including lobules IV and V of anterior lobe, Crus I, Crus II, and paramedian lobule also showed increases in activity for delay conditioning but not for trace conditioning. Delay and trace eyeblink conditioning both resulted in increased metabolic activity within the cerebellum but delay conditioning resulted in more widespread cerebellar cortical activation.  相似文献   

12.
Rats administered the cannabinoid agonist WIN55,212-2 or the antagonist SR141716A exhibit marked deficits during acquisition of delay eyeblink conditioning, as noted by Steinmetz and Freeman in an earlier study. However, the effects of these drugs on retention and extinction of eyeblink conditioning have not been assessed. The present study examined the effects of WIN55,212-2 and SR141716A on retention and extinction of delay eyeblink conditioning in rats. Rats were given acquisition training for five daily sessions followed by one session of retention training with subcutaneous administration of 3 mg/kg of WIN55,212-2 or 5 mg/kg of SR141716A and an additional session with the vehicle. Two sessions of extinction training were then given with WIN55,212-2, SR141716A, or vehicle. Retention and extinction were impaired by WIN55,212-2, whereas SR141716A produced no deficits. The extinction deficit in rats given WIN55,212-2 was observed only during the first session, suggesting a specific impairment in short-term plasticity mechanisms. The current results and previous findings indicate that the cannabinoid system modulates cerebellar contributions to acquisition, retention, and extinction of eyeblink conditioning.  相似文献   

13.
We tested the proposal that trace and delay eyeblink conditioning are fundamentally different kinds of learning. Strings of one, two, three, or four trials with the conditioned stimulus (CS) alone and strings of one, two, three, or four trials with paired presentations of both the CS and the unconditioned stimulus (US) occurred in such a way that the probability of a US was independent of string length. Before each trial, participants predicted the likelihood of the US on the next trial. During both delay ( n =20) and trace ( n = 18) conditioning, participants exhibited high expectation of the US following strings of CS-alone trials and low expectation of the US following strings of CS-US trials a phenomenon known as the gambler's fallacy. During delay conditioning, conditioned responses (CRs) were not influenced by expectancy but by the associative strength of the CS and US. Thus, CR probability was high following a string of CS-US trials and low following a string of CS-alone trials. The results for trace conditioning were opposite. CR probability was high when expectancy of the US was high and low when expectancy of the US was low. The results show that trace and delay eyeblink conditioning are fundamentally different phenomena. We consider how the findings can be understood in terms of the declarative and nondeclarative memory systems that support eyeblink classical conditioning.  相似文献   

14.
It has been proposed that the medial prefrontal cortex (mPFC) is not necessary for delay eyeblink conditioning (DEC). Here, we investigated the involvement of the mPFC in DEC with a soft or loud tone as the conditioned stimulus (CS) by using electrolytic lesions or muscimol inactivation of guinea pig mPFC. Interestingly, when a soft tone was used as a CS, electrolytic lesions of the mPFC significantly retarded acquisition of the conditioned response (CR), and muscimol infusions into mPFC distinctly inhibited the acquisition and expression of CR, but had no significant effect on consolidation of well-learned CR. In contrast, both electrolytic lesions and muscimol inactivation of mPFC produced no significant deficits in the CR when a loud tone was used as the CS, or in the unconditioned response (UR) when a soft or loud tone was used as the CS. These results demonstrate that the mPFC is essential for the DEC with the soft tone CS but not for the DEC with the loud tone CS.  相似文献   

15.
Ovarian hormones modulate acquisition processes involved in classical conditioning. Although progesterone has been indirectly implicated, its role in classical conditioning of the eyeblink response has not been directly investigated. We assessed the effects of daily dosing of progesterone or medroxyprogesterone (MPA), a non-metabolized synthetic progestin, upon the acquisition of a classically conditioned eyeblink response in ovariectomized (OVX) female rats. Rats were dosed 4h prior to each training session with 0.1 or 1.5 mg/kg of either of these hormones or sesame oil. A delay conditioning paradigm was employed using a 500 ms conditioned stimulus coterminating with a 10 ms 10 V unconditioned stimulus. At the low dose, progesterone and MPA rats did differ from each other, with MPA-treated rats learning slower, but neither group differed from OVX-oil or Sham-oil controls. No group differences in acquisition were observed at the higher dose. During extinction trials, high-dose MPA-treatment and OVX-oil groups extinguished quicker than the high-dose progesterone-treated group. In addition, unconditional response (UR) amplitudes were lower in all OVX groups, regardless of hormone or oil treatment, compared to the sham-oil group. Since MPA did not affect extinction, it is likely the slower extinction in the progesterone-treated rats is due to a metabolite of progesterone. Corticosterone is discussed as a likely candidate for such a role. In addition, we found chronic absence of ovarian hormones decreased UR amplitudes, although differences in UR amplitudes were not associated with changes in the acquisition process. These results are discussed with respect to differences in the hormonal effects upon acquisition versus extinction processes and how these data may explain reports of learning differences in women based on oral contraceptive usage.  相似文献   

16.
A common cellular alteration, reduced post-burst afterhyperpolarization (AHP) in CA1 neurons, is associated with acquisition of the hippocampus-dependent tasks trace eyeblink conditioning and the Morris water maze. As a similar increase in excitability is correlated with these two learning paradigms, we sought to determine the interactive behavioral effects of training animals on both tasks by using either a consecutive or simultaneous training design. In the consecutive design, animals were trained first on either the trace eyeblink conditioning task for six sessions, followed by training on the water maze task for six sessions, or vice versa. The simultaneous design consisted of six or 11 training days; animals received one session/day of both trace eyeblink conditioning and water maze training. Separate groups were used for consecutive and simultaneous training. Animals trained on both tasks simultaneously were significantly facilitated in their ability to acquire the trace eyeblink conditioning task; no effect of simultaneous training was seen on the water maze task. No effect was seen on acquisition for either task when using the consecutive training design. Taken together, these findings provide insight into how the hippocampus processes information when animals learn multiple hippocampus-dependent tasks.  相似文献   

17.
Three converging lines of evidence have suggested that cerebellar abnormality is implicated in developmental language and literacy problems. First, some brain imaging studies have linked abnormalities in cerebellar grey matter to dyslexia and specific language impairment (SLI). Second, theoretical accounts of both dyslexia and SLI have postulated impairments of procedural learning and automatisation of skills, functions that are known to be mediated by the cerebellum. Third, motor learning has been shown to be abnormal in some studies of both disorders. We assessed the integrity of face related regions of the cerebellum using Pavlovian eyeblink conditioning in 7–11 year-old children with SLI. We found no relationship between oral language skills or literacy skills with either delay or trace conditioning in the children. We conclude that this elementary form of associative learning is intact in children with impaired language or literacy development.  相似文献   

18.
Renewal of an extinguished conditioned response has been demonstrated in humans and in animals using various types of procedures, except renewal of motor learning such as eyeblink conditioning. We tested renewal of delay and trace eyeblink conditioning in a virtual environment in an ABA design. Following acquisition in one context (A, e.g., an airport) and extinction in a different context (B, e.g., a city), tests for renewal took place in the acquisition (A) and extinction context (B), in a counterbalanced order. Results showed renewal of the extinguished conditioned response in the delay but not trace condition.  相似文献   

19.
In delay eyeblink conditioning, the CS overlaps with the US and only a brainstem-cerebellar circuit is necessary for learning. In trace eyeblink conditioning, the CS ends before the US is delivered and several forebrain structures, including the hippocampus, are required for learning, in addition to a brainstem-cerebellar circuit. The interstimulus interval (ISI) between CS onset and US onset is perhaps the most important factor in classical conditioning, but studies comparing delay and trace conditioning have typically not matched these procedures in this crucial factor, so it is often difficult to determine whether results are due to differences between delay and trace or to differences in ISI. In the current study, we employed a 580-ms CS-US interval for both delay and trace conditioning and compared hippocampal CA1 activity and cerebellar interpositus nucleus activity in order to determine whether a unique signature of trace conditioning exists in patterns of single-unit activity in either structure. Long-Evans rats were chronically implanted in either CA1 or interpositus with microwire electrodes and underwent either delay eyeblink conditioning, or trace eyeblink conditioning with a 300-ms trace period between CS offset and US onset. On trials with a CR in delay conditioning, CA1 pyramidal cells showed increases in activation (relative to a pre-CS baseline) during the CS-US period in sessions 1-4 that was attenuated by sessions 5-6. In contrast, on trials with a CR in trace conditioning, CA1 pyramidal cells did not show increases in activation during the CS-US period until sessions 5-6. In sessions 5-6, increases in activation were present only to the CS and not during the trace period. For rats with interpositus electrodes, activation of interpositus neurons on CR trials was present in all sessions in both delay and trace conditioning. However, activation was greater in trace compared to delay conditioning in the first half of the CS-US interval (during the trace CS) during early sessions of conditioning and, in later sessions of conditioning, activation was greater in the second half of the CS-US interval (during the trace interval). These results suggest that the pattern of hippocampal activation that differentiates trace from delay eyeblink conditioning is a slow buildup of activation to the CS, possibly representing encoding of CS duration or discrimination of the CS from the background context. Interpositus nucleus neurons show strong modeling of the eyeblink CR regardless of paradigm but show a changing pattern across conditioning that may be due to the necessary contributions of forebrain processing to trace conditioning.  相似文献   

20.
The current study examined the effects of globally depleting Purkinje cells in the cerebellar cortex with the immunotoxin OX7-saporin on acquisition and extinction of delay eyeblink conditioning in rats. Rats were given OX7-saporin or saline 2 wk before the start of eyeblink conditioning. The rats that reached a performance criterion of two consecutive days with 80% or greater conditioned responses were given 5 d of extinction training followed by 2 d of reacquisition training. Rats that received infusions of OX7-saporin had 77.2%-97.9% Purkinje cell loss and exhibited impaired acquisition and extinction. The amount of Purkinje cell loss was correlated with the magnitude of the acquisition and extinction impairments. The highest correlations between Purkinje cell number and the rate of acquisition were in lobule HVI and the anterior lobe. The highest negative correlation between Purkinje cell number and the percentage of conditioned responses during extinction was in the anterior lobe. The results indicate that cerebellar Purkinje cells, particularly in the anterior lobe and lobule HVI, play significant roles in acquisition and extinction of eyeblink conditioning.  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