European Journal of Neuroscience, Vol. 20, pp. 791–802, 2004

ª Federation of European Neuroscience Societies

Reward expectation, orientation of attention and locus coeruleus-medial frontal cortex interplay during learning Sebastien Bouret and Susan J. Sara Neuromodulation et Processus Mne´siques, CNRS UMR 7102, Universite´ Pierre & Marie Curie, 9 quai St. Bernard, 75005 Paris, France Keywords: incentive, motivation, neuromodulation, noradrenaline, prefrontal cortex, rat

Abstract Regulation of attention and promotion of behavioural flexibility are functions attributed to both the noradrenergic nucleus locus coeruleus (LC) and the prefrontal cortex (PFC). The PFC receives a large innervation from LC and small changes in catecholaminergic activity in PFC profoundly affect cognitive function. It is crucial to the understanding of learning-related plasticity, that the cognitive context driving LC neurons be determined and the relation to activity in PFC be elucidated. To this end simultaneous recordings were made from LC and prelimbic cortex (PL) during an odour-reward association task in the rat. Neuronal activity related to orientation of attention, reward predictability, reward itself, and changes in stimulus reinforcement contingencies, was measured. All LC neurons and a significant proportion of PL neurons were engaged during several aspects of a Go ⁄ NoGo task, especially after the signal for trial onset and CS+ presentation. LC activation was, however, more tightly aligned to the behavioural response than to the CS+ 22% of PL neurons were activated during the response-reward delay. This suggests that the activity of both these structures is related to reward anticipation. Finally, LC neurons exhibited rapid plasticity when the reward-contingency was modified. Within-trial response latencies were always shorter in LC than in PL and between-trial response adaptation in LC preceded that in PL by many trials. Identifying such temporal relationships is an essential step toward understanding how neuromodulatory inputs to forebrain networks might promote or permit experience-dependent plasticity in behavioural situations.

Introduction Research over the past two decades has established that noradrenaline (NA) modulates sensory information processing (Waterhouse & Woodward, 1980; Bouret & Sara, 2002; Lecas, 2004) and promotes synaptic plasticity in forebrain regions (Harley et al., 1983; Walling & Harley, 2004). Evidence comes mainly from in vivo and in vitro electrophysiological studies and has led many investigators to assign an important role to the noradrenergic system in regulation of attention, learning and memory (reviewed in Berridge & Waterhouse, 2003). To further support this notion, single unit recordings in behaving primates and rats have shown that noradrenergic neurons of the locus coeruleus (LC) have distinct stimulus-related activity. Novel sensory stimuli and changes in their predictive value elicit phasic LC responses, suggesting a role for LC in regulating attention and promoting behavioural flexibility (Sara & Segal, 1991; Vankov et al., 1995; Aston-Jones et al., 1997). Conclusions concerning the essential role of the LC neuromodulatory system in cognitive processes are presently based on these two distinct areas of research, one showing postsynaptic modulatory influence of and the other elucidating the behavioural context driving noradrenergic neurons. These conclusions would be greatly strengthened by a precise description of the temporal relationship between LC activation and target forebrain activity in well-defined cognitive situations. Lesion studies in rodents have suggested that, like LC, the prelimbic (PL) area of the medial frontal cortex (mFCx), is involved in attention

Correspondence: Dr Susan J. Sara, as above. E-mail: [email protected] Received 3 March 2004, revised 26 May 2004, accepted 28 May 2004

doi:10.1111/j.1460-9568.2004.03526.x

and behavioural flexibility (Delatour & Gisquet-Verrier, 2000; Birrell & Brown, 2000; Passetti et al., 2002). Thus, PL would be a likely candidate region for studying neuronal activity in parallel with that of LC in controlled behavioural conditions. While much is known about dopaminergic influences in PFC, the noradrenaline (NA) system has been much less studied (Arnsten, 1997), even though the LC projection to this region is substantial (Florin-Lechner et al., 1996). Available evidence from electrophysiological studies in rat suggests that NA has an inhibitory effect in PFC, but evoked activity is spared thereby improving the signal-to-noise ratio (Mantz et al., 1988). Similar conclusions have been drawn from studies of the effects of NA on PFC neuronal responding in primates performing a delayed response task (Sawaguchi & Kikuchi, 1998). Nevertheless, the relationship between LC activation and PFC activity emerging from complex network interactions, remains unknown. Against this background, one aim of the present study is to examine neuronal activity in LC and PL in well-defined behavioural conditions to determine the similarities and differences in response patterns and timing. We recorded simultaneously in PL and LC while the rat performed a Go ⁄ NoGo olfactory discrimination task. The protocol included a preparatory stimulus, light on, to elicit attention at the beginning of the trial. The differential Go ⁄ NoGo procedure enabled us to look at neuronal activity during sensory processing as well as during goal-directed and error-related behaviour. A delay between the correct behavioural response to CS+ and reward delivery is used to examine reward prediction activity. LC and PL unit activity is characterized during each step of this task. Recording over several daily sessions allowed evaluation of neuronal activity during behavioural adaptation to changes in odour-reward contingency

792 S. Bouret and S. J. Sara (new discriminanda, reversals, and extinction). Most importantly, we recorded simultaneously from the neuromodulatory LC cells projecting to the entire forebrain, and from a target region receiving multiple and diverse inputs, during behavioural situations requiring rapid adaptation to changing reward contingencies. This should significantly advance our understanding of how neuromodulatory influences promote or permit neuronal adaptation to changing cognitive demands in forebrain integrative networks.

Experimental procedures Animals Electrophysiological recordings were taken from seven male Sprague– Dawley rats obtained from IFFA Credo (L’Arbresle, France). The rats, weighing 280–380 g at the time of surgery, were housed for two weeks before the experiment in a temperature controlled vivarium on a 12-h light : 12-h dark cycle. They were weighed and handled regularly and had free access to water. Access to food was regimented throughout training and recording sessions and rats were maintained at approximately 90% of their freely feeding weight. Food was available ad libitum for the week that followed surgery. Surgical procedures and experimental techniques were carried out according to the 1986 European Communities Council Directive and Ministere de l’Agriculture et de la Foret, Commission Nationale de l’Experimentation Animal decree 87848.

Behaviour Behavioural apparatus The training took place in a 26 · 26 · 60 cm plexiglas chamber with a grid floor. The cage had a small window (reward port) to which was fixed an automated dipper that delivered a drop of chocolate milk. A photoelectric cell monitored Go responses. A light bulb was fixed on the opposite side of the box. Odours were delivered via a compressed air system; a pressure controlling device delivered a constant air flow (500 mL ⁄ min). Computer-controlled valves gated air flow through one of three tubes (1.5 mL) containing cotton impregnated with distilled water (neutral odour), CS+ or CS– odour. Flexible tubes extending from each container were tightly held in a common output tube, inserted in the odour port, located 1 cm from the side of the reward port. For olfactory stimulation, the valve gating air flow to the tube containing the odour was opened while the valve controlling the neutral air path was closed. Odours were: vanilla, almond, strawberry, lemon, mint, orange, all of which had been used for other behavioural studies (Tronel & Sara, 2002). Naive rats do not show a preference for any of these odours. The entire chamber was enclosed in a sound attenuated, electrically shielded box fitted with a transparent window through which the rat could be observed. A small extractor fan limited odour accumulation and provided background noise. All the devices were controlled by a PC computer via a Cambridge Electronic Device CED 1401 plus interface (Cambridge, England). Behavioural task Go and NoGo trials were presented in a pseudo-random order (no more than two consecutive trials of the same type) with a variable interval of 16–20 s. A constant flow of neutral air was maintained between trials. Each trial started with illumination of the light bulb (preparatory signal) that remained lit for the duration of the trial. After a variable delay of 1.5–2.5 s, one of the two odours was delivered. Animals were trained to sniff the odour port after the

preparatory signal before either odour was delivered. The odour was presented for 700 ms, but from 600 ms after odour onset, any Go response, monitored by the photocell beam, could lead to the trial outcome and terminate the odour delivery. The decision period, during which the animal had to respond after CS+ or withhold responding after CS–, was 1.76 s. Failure to respond within 1.76 s was counted as an omission error. If the animal responded in time to CS–, a commission error was counted. For either correct Go responses or commission errors, a 500 ms delay was introduced between the response and the outcome. Because the minimum interval between odour onset and the beginning of the 500 ms-delay period was fixed set at 600 ms, and in most cases the behavioural latency was below 600 ms (mean latency, 573 ms), the actual duration of the interval between the Go response as measured by the photocell breaking and the reward delivery was variable, usually longer than 500 ms. Trial termination was indicated by turning off the light at the end of the decision period if the animal did not respond to the odour (omission error after CS+ or correct NoGo after CS–). Correct Go responses after CS+ were automatically rewarded by a drop of chocolate milk delivered via the dipper that remained up for 2 s, after which the light was turned off. Incorrect Go responses to the CS– odour (commission errors) were signalled by the light being turned off at the end of the postresponse 500 ms delay, and punished by an extended intertrial interval (a 30 s timeout in addition to the variable 18 s intertrial interval). The first two animals were trained on a slightly different version of the task: odour duration was longer (2 s) and could be interrupted by a Go response 1 s after onset. The decision period, during which the animal could choose between a Go and a NoGo response extended to 4 s after CS onset. Because the animal could respond with a latency as short as 600–700 ms, odour duration and decision period were shortened, and the next five rats were run with the previously described protocol. Although rats trained with the first protocol showed a longer mean behavioural latency (1.4 s) than rats run with the second version (573 ms), no other difference was observed nor did we find any difference in neuronal activity in the PL between these two protocols. Hence, results will be presented together. No reliable LC recording was obtained from the first two rats. Training Animals were initially habituated to the apparatus and trained to drink from the dipper. They were then progressively shaped to sniff the odour port after the light and to Go and collect the reward after the CS+ odour delivery. They were trained until they met a criterion of > 90% correct response for both CS+ and CS– odours, before electrodes were implanted. This training phase lasted from four to six days (130 trials a day). Subsequent daily recording sessions comprised 150–300 trials. Learning situations For four rats from which reliable LC recordings were obtained, several new learning situations were imposed. The first was an intradimensional shift (IDS). During each IDS session, a new set of odours was introduced and the animal discovered the stimulus-reward contingency of the novel odours by trial and error. Rats were trained to a criterion of > 90% correct response with the new odours. For two animals, several IDS sessions were repeated before an extinction protocol. For extinction the dipper was disconnected in the middle of the session, without any break or information for the animal, and reward was unavailable until the animal showed a clear behavioural

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

Locus coeruleus-frontal cortex interplay 793 extinction (no response to CS+ for ten successive trials). Once extinction was established, the dipper was reconnected. During this relearning phase, correct Go responses were rewarded and rats rapidly returned to asymptotic performance. After several extinction sessions (one for each pair of odours), these rats were submitted to reversal sessions. The odour-reward contingency was reversed in the middle of the session and rats were trained until they expressed a clear discrimination (> 80% correct responses). A complete behavioural reversal (90% correct responses) was never observed during the first reversal session, and an additional session was necessary to reach asymptotic performance. For subsequent reversal sessions, one of the previous sets of odours was used, and the animal showed perfect retention. In contrast to the initial reversal training, subsequent reversals were learned within one session.

Electrophysiology Single-unit recordings were made through etched tungsten microelectrodes, insulated with epoxylite; standard medium tip, 1–3 lm; impedance 1–3 MW (FHC, Inc, Bowdoinham, Me) glued to a movable microdrive. For PL, differential recordings between two microelectrodes glued together on the same micromanipulator allowed recording of 2–4 units simultaneously with minimum muscle or movement artefacts (200–500 lm between tips). For LC recordings, an independent fixed reference electrode was implanted in the midbrain. An uninsulated stainless steel wire wrapped around a skull screw served as ground.

the connector containing the ground wire, reference electrode and leads for microelectrode connection. Animals were returned to their home cage for seven days of recovery. Chronic recording procedure The animal was connected to a cable containing field effect transistors. The signal was filtered (300–3000 Hz band pass), amplified (·10K) (A-M Systems model 1700) and displayed on an oscilloscope and audio monitor. The signal was digitalized at 33 kHz and wave forms were discriminated on line by a template matching algorithm using the Cambridge Electronic Design (CED, Cambridge, England) CED1401 digital converter and Spike2 software (CED). Data were stored on a PC computer for further off line analysis. Electrodes were advanced in increments of 40 lm until a well differentiated unit was encountered. The PL electrode assembly was advanced at the beginning of each session, whereas the LC electrode was left in place as long as a typical unit or multiunit LC recording was observed. Histology After completion of the experiments, animals were deeply anaesthetized and perfused transcardially with saline and 4% formalin. After one week postfixation in formalin, brains were extracted and sectioned at 60 microns. Sections were mounted and stained with cresyl-violet. Only recordings from animals for which correct electrode location was confirmed after examination of histological sections are reported (n ¼ 7). Figure 1 shows representative histological sections.

Surgery

Data analysis

Animals were anaesthetized with pentobarbital (60 mg ⁄ kg, supplemented as necessary). They were fixed in a stereotaxic frame with the nose angled down (16). The skull was exposed and holes were drilled for anchoring skull screws and for the LC reference electrode. The two movable microdrive assemblies were successively advanced into corresponding holes drilled above PL (3.2 mm anterior to bregma, 0.6 mm from midline) and LC (4 mm posterior to lambda, 1.15 mm lateral to midline) under electrophysiological control. The LC nucleus, located 5–6 mm below the cerebellar surface, was identified by the typical activity of these neurons [broad spikes (> 0.7 ms), low firing rate (< 2 Hz) (see Shinba et al., 2000 fur further details)]. The PL electrode was positioned 2 mm below the brain surface and the LC electrode at 100–200 microns above the LC nucleus. The two microdrive assemblies were fixed to the skull with dental cement, along with

Single units were isolated wherever possible, using the Spike2 software. Further spike sorting was made off-line. If the spikes were not clearly separable, the file was treated as a multiunit recording. Post stimulus time histograms (PSTHs) and raster displays were generated for neuronal activity 500 ms before and 500 ms after light onset, using 20 ms bins. The mean and standard deviation (SD) of neuronal firing activity was calculated for the 500 ms prestimulation baseline. A firing rate increase to 2SDs above the mean of the base line, sustained over at least three bins, was considered an excitatory response. A decrease to 2SDs below the mean was considered an inhibitory response. Response latencies were thus calculated for each unit or multiunit record. A similar approach was used to calculate the latency of the response to odours. Many cells showed a change in activity before odour onset. Spike counts during response windows were computed for successive trials,

Fig. 1. Histology. (Left) Histological section showing an electrode track passing through the cerebellum and terminating in the medial part of LC (arrow). On the contralateral side, the LC is clearly visible, located below the fourth ventricle (IV) and medial to the dorsal tegmental area (DT). (Right) Histological section in the frontal region (3.2 mm anterior to bregma). The electrode track is clearly visible in the deep layers of the medial frontal cortex. The track terminates in the PL area (arrow). ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

794 S. Bouret and S. J. Sara for different events of the task. Several windows were considered: a 500 ms preodour window, ending with odour onset; a 500 ms odour window, starting 50 ms after odour onset, a 500 ms interval window, during the response-outcome interval, ending with the outcome (reward or light off for commission errors) and a 500 ms window starting with the trial outcome (reward, light off for commission errors or correct NoGo or end of reward delivery period). Spike counts during these windows were compared to spike counts during an equivalent baseline window, i.e. ending with light onset, before the beginning of the trial. Spike counts in the response and baseline windows were compared with a non-parametric Wilcoxson signed rank test, as most spike counts were not normally distributed. Nonparametric Friedman anova followed by Newman–Keuls analogue test as posthoc analysis, was used for multiple comparisons. A discrimination index (DI) was calculated to normalize the LC differential response to CS+ and CS–. This was obtained by dividing the difference between response to each odour by the sum of these values, yielding a value between +1 and )1. A response index was similarly calculated by dividing the difference between response and baseline spike counts by the sum of these values. For population analyses, firing rates were averaged across cells and compared by paired t-tests. Firing rate ratios between baseline and response were used for LC population analysis, because of the high variability in baseline activity between multiple and single unit recordings. Mean response ratios were compared to unity using one sample t-tests. In addition, LC activity alignments on CS+ onset or behavioural response (measured by the photocell) were compared. PSTHs (10 ms bin) displaying LC activity 2 s before and after stimulus were constructed around CS+ onset and around corresponding Go responses, and only trials with a correct Go response were taken into account. Peak heights of these two PSTHs were compared (the peak being defined by the value of the highest bin between odour onset and behavioural response). The ratio of these two peaks was calculated for each recording session and a one sample t-test was applied to verify that this mean ratio was significantly different from one.

Results Baseline activity in PL and LC One hundred and twelve single units (SUs) were recorded from deep and superficial layers of the mFCx. Recording was restricted to the intermediate mFCx (PL) and did not include the more dorsal Cg1 or the more ventral infralimbic regions, as confirmed by histological analysis. These cells had a relatively low firing rate (mean 0.78 ± 0.08 Hz). Thirteen SUs and 36 multiple unit recordings (MUs) were obtained from the anterior portion of LC. The 13 SUs had a mean firing rate of 1.3Hz ± 0.2 Hz. The 36 MU had a mean firing rate of 2.8 Hz ± 0.3 Hz, suggesting that MUs contained 2–3 single units. Task related activity Trial initiation: light onset Behaviourally, animals orientated to the light, moved to the odour port and sniffed it actively (see Fig. 2). At light onset, a significant phasic excitation was noted in almost every LC recording [46 out of 49 (Wilcoxson, P < 0.05 for all 46 records]; all 13 SUs responded to the light (Fig. 3A). The latency was highly consistent across recordings and animals (155 ± 1.8 ms). As can be seen in Fig. 3, this period of excitation lasted approximately 100 ms and was sometimes followed by a short-lasting (80–100 ms) phase of inhibition (especially clear for SU recordings). It should be noted that this response consisted of only a few spikes (n ¼ 1– 3 spikes SUs) well aligned to light onset. Thus, the response was not readily discernible if LC activity was triggered on any event other than the light. Considering the entire population of 49 LC recordings, there was an overall 33.5 ± 2.8% increase in firing rate from baseline during the 500 ms window after light onset (significantly different from 0, one sample t-test, t48 ¼ 11.9, P < 0.0001). Independently of the postexcitation inhibition, 32 out of 49 LC recordings showed a significant decrease in activity during the 500 ms preodour period (Wilcoxson, P < 0.05 for each of the 32 recordings),

Fig. 2. Behaviour. (A) Schematic representation of the experiment. During the intertrial interval, the light is off and the dipper is out of reach. A constant stream of air flows through the odour port (Od port). (B) At trial initiation (top left), the lamp lights, the rat moves to the odour port, sniffs and waits for odour delivery. Welltrained rats disengage for CS–. For CS+ (top right), the rat quickly moves to the reward port and interrupts the photo-electric cell (bottom right). It has to wait at least 500 ms after the response has been registered for reward delivery via the dipper (bottom left, filled circle). The dipper is returned to its initial position after 2 s, and the light is extinguished. ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

Locus coeruleus-frontal cortex interplay 795 Fig. 3. Response to light in LC and PL. (A) Left, PSTH (10 ms bin) and raster display for LC SU activity (baseline firing rate across trials 3.1 ± 0.1 Hz) around light onset. Light onset elicits a clear phasic excitation (42% increase in firing rate, Wilcoxson: z ¼ )3.6, P ¼ 0.0003) with a latency of 130 ms. Note the short lasting phase of postexcitation inhibition (arrow), particularly visible on the raster. Right, PSTH of the same recording around light onset with a wider time scale (50 ms bin). Rasters are sorted by increasing latencies between light onset and odour onset (grey circles). Note, in addition to the phasic excitatory response, a significant late inhibition (81% from baseline firing rate, Wilcoxson, z ¼ )7.13, P < 0.0001), particularly visible right before odour onset, when the animal was actively sniffing the odour port. (B) Left, activity of a PL unit (1.8 ± 0.2 Hz baseline firing rate) around light onset (10 ms bin). This neuron shows a clear inhibition 250 ms after light onset (58% from baseline firing rate, Wilcoxson, z ¼ )3.9, P < 0.0001). Circles indicate the first photocell interruption (Go response) after light onset. Right, raster constructed around light onset for the same unit. Vertical lines indicate light onset and average light offset. Trials are sorted by increasing latencies between light onset and first photocell interruption (circles). Small grey bars indicate odour onset. Note the tonic inhibition that lasts until the end of each trial, without being affected by odour delivery or behavioural response. (C) Raster display around light onset (vertical line) for a LC MU (2.2 ± 0.4 baseline firing rate, left) recorded simultaneously with a PL unit (1.24 ± 0.2 Hz baseline firing rate, right) during an extinction session. Both records show a significant change in response magnitude across this session (Friedman anova, v2(5) ¼ 16.5 P ¼ 0.0006 for PL, v2(5) ¼ 18.5, P ¼ 0.002 for LC. During asymptotic performance (trials 1–90), LC neurons show a short latency phasic excitatory response (63% increase from baseline, Newman–Keuls analogue, P < 0.05) while the cortical unit shows a tonic inhibition (71% from baseline firing rate Newman–Keuls analogue, P < 0.05). During extinction (trials 91–180, shaded area), both LC and PL stop responding to the light (no difference from baseline, Newman–Keuls analogue, P > 0.05). Bar indicates period of behavioural extinction (no response to CS+). Responses to light return to pre-extinction values when the reward is restored during the following relearning phase (trials 181–235) (LC, 46% increase from baseline; FC, 80% decrease from baseline firing rate; Newman–Keuls analogue P < 0.05).This phenomenon was observed for all LC recordings tested in this situation (n ¼ 11).

rons, there was a 47 ± 2% decrease in firing rate after light onset. The latency was relatively long (248 ± 15 ms) with the difference in latencies between LC and PL neurons significant (t36,32 ¼ 6.1 P < 0.0001). For ten PL cells, the activity returned to baseline before odour onset (phasic response). For the remaining 32 neurons, the inhibition lasted at least until the end of the odour-on period (tonic response). The change in neuronal activity is not solely related to modification of locomotor activity. Neuronal inhibition was observed both when the animal orientated and approached the reward port before moving to the odour port and when it was standing in front of the odour port (Fig. 3B). The light-evoked activity in both LC and PL disappeared during the course of extinction, when the animal no longer systematically oriented to the stimuli (Fig. 3C). When reward was restored, a rapid increase in behavioural reactivity was observed, in parallel with the phasic excitatory responses in LC and the tonic inhibitory responses in PL. This light evoked activity in all LC and PL recordings remained stable during Intra-dimensional shift (IDS) and reversal sessions, even before the odour–reward association was discovered, as long as the animal kept performing the task. Cue: odour delivery when the animal was actively sniffing the odour port (Fig. 3A, right). At a population level, a 34% ± 0.04% decrease in firing rate from baseline was observed during the preodour period (significantly different from 0, one sample t-test, t48 ¼ 7.7, P < 0.0001). In PL, 42 neurons (38%) showed a significant change in activity after light onset (Wilcoxson, P < 0.05 for all 42 records), all but one showing an inhibitory response (Fig. 3B). Considering these 42 neu-

After CS–, animals sniffed the odour port and disengaged. After CS+ , animals moved rapidly to the reward port, interrupting the photocell beam. All LC recordings showed a differential response to olfactory cues, as shown in Fig. 4. During CS–, the activity of those LC cells gradually decreased before odours, slowly returned to baseline (latency 250–600 ms). During CS+, a late excitation was consistently observed (mean latency 257 ± 14 ms), the mean firing rate during

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

796 S. Bouret and S. J. Sara

Fig. 4. Differential response to odour cues in LC. (Top) PSTH (50 ms bins) and raster for a MU LC recording (3.2 ± 0.3 Hz baseline firing rate) triggered on odour for CS– (left) and CS+ (right). On the raster, trials are sorted by increasing behavioural response latencies. These cells showed a significant inhibition before odours (75% decrease from baseline firing rate, Wilcoxson: z ¼ )6.3, P < 0.0001), when the animal was actively sniffing the odour port after light onset. During CS–, the activity returns to baseline 330 ms after odour onset. During CS+, there is a significant excitation (latency ¼ 230 ms, 101% increase from baseline firing rate, Wilcoxson: z ¼ )5.5, P < 0.0001), before the animal moves to the reward port and interrupts the photocell beam (circles, nose-pokes). Note that the activity shows a further decrease 1 s after odour onset, when the animal is consuming the reward (delivered 600 ms after behavioural response). (Bottom) Raster and corresponding PSTHs (10 ms bins) show the activity of a LC unit around odours trials are sorted by increasing behavioural response latencies. The animal showed a perfect performance for CS+ trials (no omission). Left, activity aligned on CS+ onset (line), circles indicate photocell interruption. Middle, same activity aligned on the behavioural response (line), circles now indicate CS+ odour onset. Right, activity aligned on incorrect response (commission errors, line) after CS– (circles). Comparison of the left and the middle display shows that LC excitation to the CS+ odour is actually more tightly related to the behavioural response. All LC recordings showed a similar phenomenon. The right display shows that this LC activation is not motor related: it is much weaker or absent when the animal shows a behavioural response to CS– that does not predict reward delivery.

CS+ showed a 51% ± 8% increase from baseline (significantly different from 0, one sample t-test: t48 ¼ 6.7, P < 0.0001). For all LC recordings obtained, the firing rate was significantly higher during CS+ than during CS– (Wilcoxson, P < 0.05 for each of the 49 LC recordings). A discrimination index was calculated (see Experimental procedures) for each cell, and the mean for the population (0.29 ± 0.02) was significantly different from 0 (one sample t-test, t48 ¼ 6.7, P < 0.0001). This phasic excitatory response was further examined. PSTHs constructed around CS+ onset or around the behavioural response were compared (Fig. 4, bottom). LC activation was more tightly aligned to the behavioural response, measured by photocell interruption, than to odour onset. In order to quantify this phenomenon, peak heights of these two PSTHs (constructed with the same total number of spikes) were compared. A significant difference in peak height was noted for all LC recordings, reflecting a better aligning of LC activation with behavioural response than to CS+ (mean ratio of peak heights of response triggered PSTHs vs. odour triggered PSTHs ¼ 1.27 ± 0.04, significantly different from 1; t48 ¼ 7.4, P < 0.0001). Comparison of peak latency from odour onset (269 ± 12 ms) and before behavioural response (212 ± 7 ms) also showed a clear difference (t48 ¼ 3.5; P ¼ 0.0009), with the excitatory response occurring closer in time to the behavioural response than to the odour onset (Fig. 4 bottom). Moreover, when LC activity was examined during the few CS+ trials when the animals did not show a

Go response (omission), no activation was observed. In order to see whether LC activation was related to locomotor activity, LC activity was examined during CS– trials when animals made a commission error, i.e. when they made an incorrect Go response. Here, no LC activation was noted before the behavioural response, and no difference between odour- and behavioural response-triggered PSTHs was observed. LC activation to CS+ is therefore not simply related to locomotor activity. Taken together, these data show that the late excitation to CS+ in LC neurons is tightly related to the rewarddirected behavioural response. In PL, few neurons (n ¼ 7, 6%) showed a response to odours and six out of seven were inhibitory responses. Response-reward interval During the interval between behavioural response to CS+ and reward delivery, animals stayed in the reward port until the end of reward delivery period. For commission errors outside of reversal sessions, both strength and duration of behavioural responses to CS– were weaker than responses to CS+ , as measured by response latency. In PL, 25 cells (22%) showed a large modification of their firing rate during this period. Most neurons (19 out of 25) showing a response during the interval were excited (Fig. 5)(Wilcoxson, P < 0.05 for each of the 19 neurons). For these 19 units, the average firing rate increased from 0.63 ± 0.2 Hz to 1.67 ± 0.4 Hz, t18 ¼ )3.6, P ¼ 0.002). The other six neurons were inhibited during the response-

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

Locus coeruleus-frontal cortex interplay 797

Fig. 5. PL activity during the response-reward interval. PSTH (50 ms bin) and raster display for a PL unit (0.28 ± 0.08 Hz baseline firing rate) showing an excitation during the response-reward interval. Trials are sorted by increasing response-outcome latencies. Twenty-two per cent of PL units showed a response during this interval. (Left) The activity is aligned on reward delivery (vertical line) after correct behavioural response (circle). The response is restricted to the interval between the behavioural response and the reward (214% increase in firing rate from baseline, Wilcoxson: z ¼ )3.2, P ¼ 0.002). (Right) No excitatory response is observed after the few comission errors (vertical line indicates light off at end of incorrect response trials, circle indicates incorrect photocell interruption). Hence, the activity during the delay is not related to the motor response (Go) but to the anticipated reward delivery.

reward interval (average firing rate decreasing from 0.82 Hz ± 0.13 to 0.43 Hz ± 0.08, t5 ¼ 4, P ¼ 0.009). It should be noted that this interval activity was never observed after an incorrect Go response to CS– (Fig. 5), suggesting that it is related to the anticipation of the forthcoming reward delivery and not simply to the behavioural response. Reward delivery During asymptotic performance sessions, LC neurons showed no phasic response to trial outcome (reward delivery, errors, end of NoGo trial) but showed a decrease in activity during reward consumption (see Fig. 4 top). In PL, 23 neurons (21%) showed a response related to reward, occurring either at reward delivery (n ¼ 18) or termination (n ¼ 5). Eight of these cells responded both before and after reward delivery.

Learning related changes There were three distinct learning situations in which stimulusreinforcement contingencies were changed, requiring behavioural adaptation: (i) new odours (intra-dimensional shift, IDS); (ii) extinction and (iii) reversal. These changes elicited rapid adaptive responses in LC, that were strikingly homogeneous across cells. Behaviour Six IDS sessions were run with a new pair of stimuli being introduced during a session of asymptotic performance. In early IDS trials, although animals initially actively sniffed the novel stimuli, they did not respond to either new odour for several trials (4.3 ± 1.4). Then, they showed a systematic Go response to both odours for an average of 8.3 ± 2.3 trials, before showing a behavioural discrimination, by only responding to the new CS+. Eight extinction sessions were completed. A progressive increase in behavioural response latency always preceded complete behavioural

Fig. 6. LC discrimination precedes behavioural discrimination during IDS. Mean behavioural performance (n ¼ 6 sessions) and mean response index (RI, n ¼ 6 LC records) for consecutive blocks of ten trials (five of each type). A significant change in neuronal [Friedman anova, v2(9) ¼ 17, P ¼ 0.049 for LC] and behavioural (v2(4) ¼ 16, P ¼ 0.003) discrimination was observed during IDS. Before IDS, behavioural discrimination is above 80% of correct response and LC neurons clearly discriminate between CS+ and CS– (*significant difference between CS+ and CS– RI, P < 0.05 posthoc Newman–Keuls analogue test). At the beginning of IDS, both behavioural and neuronal discrimination drop significantly (  significantly different from preIDS value). Whereas LC discrimination returns to its pre-IDS level during the second block, behavioural discrimination only returns to its initial value during the third block (P < 0.05, Newman–Keuls analogue posthoc analysis). Hence, at the population level, LC discrimination precedes behavioural discrimination during IDS.

extinction. Extinction was obtained for seven out of eight sessions (within 43 ± 8 trials), whereas for one session the animal was still showing some response after more than 100 extinction trials. After complete behavioural extinction, the reward was restored. The behavioural relearning was always extremely rapid; the behavioural performance reached pre-extinction values within five trials. Five reversal sessions were run. A first correct Go response to the new CS+ odour was observed after a mean of 19 ± 7.4 trials. This was followed by persistent responding to both odours. After a large number of trials (mean 76 ± 12), response latencies to CS– increased. A complete behavioural reversal was eventually obtained after a mean of 96 ± 9 trials. A more rapid rate of reversal was seen on successive reversals. LC activity during learning During the six IDS sessions, six LC MU recordings were examined. For all six recordings, although a differential response to odours was readily observed before IDS, no initial response to novel odours was observed before the animal showed a behavioural response. All LC neurons, did however, show a marked excitation for the first few

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

798 S. Bouret and S. J. Sara

Fig. 7. Plasticity of LC response to CS+ during extinction and relearning. Raster display for the activity of a representative LC unit (2.4 ± 0.1 Hz baseline firing rate) around CS– (left) and CS+ (right) during an extinction session. Broken line indicates average behavioural response to CS+. This neuron showed a significant change in magnitude of response to CS+ during blocks of 15 trials across each step of this session (Friedman anova, v2(7) ¼ 33, P < 0.0001). Before extinction (trials 1–30), this LC neuron shows a clear excitatory response to CS+ (75% increase from baseline activity, Newman–Keuls analogue, P < 0.05). During extinction (shaded area), the excitatory response rapidly disappears (no difference between CS+ and baseline activity, Newman–Keuls analogue, P > 0.05), before the animal shows a complete behavioural extinction (vertical bar). During relearning, when the reward is restored, the cell rapidly re-acquires an excitatory response to CS+ (81% increase from baseline activity, this response magnitude being similar to its pre-extinction value, Newman–Keuls analogue, P < 0.05). The raster display on the right shows LC activity aligned on reward delivery during relearning (dotted line, average nose-poke). It shows that the cell initially responds to reward delivery (arrowhead) the first two times it is obtained, before this response rapidly shifts to the CS+ (within 2–3 trials).

reward deliveries, and this response rapidly shifted to the CS+ (within fewer than five trials), with some trials in which LC neurons responded to both CS+ and reward. As reported above, the LC activation was more aligned to the behavioural response than to the odour onset, even during the early trials. Importantly, as can be seen in Fig. 6, LC neuronal discrimination between CS+ and CS– always preceded behavioural discrimination. During the eight extinction sessions, five SU and six MU recordings were made from the LC. During the course of extinction, all LC recordings showed a very rapid decrease in response to the CS+ (within five trials), which always preceded the first signs of behavioural extinction (increased latency and omissions to CS+) (Fig. 7). No change in neuronal activity was noted at the time of usual reward delivery, when this reward was omitted. When reinforcement was restored, the first time the animal made a Go response, a strong LC activation was consistently observed at the time of reward delivery (Fig. 8, bottom right). This response to reward was maintained for a few trials (2–15), while a response to CS+ rapidly developed (within three trials). During the five reversal sessions, seven LC recordings (five MU and two SU) were obtained. For all LC recordings, a rapid extinction of the response to the previous CS+ was observed (within five trials, Fig. 8). All LC neurons were strongly excited by the first reward, following the first correct Go response to the now CS+. This excitation rapidly shifted to the CS+ odour (within 5–15 trials), but LC neurons usually responded to both CS+ and reward for a few trials, before the response to reward disappeared. LC neurons showed a selective response to CS+ at least ten trials before any behavioural discrimination was expressed (Fig. 8).

PL reward anticipation activity during learning The activity of 13 PL neurons displaying reward-anticipation activity (i.e. during the interval between behavioural response and reward delivery) were followed during these three learning situations. For all these neurons, reward anticipation activity was absent after the odour reward contingency was changed and only reappeared after the animal expressed correct behavioural discrimination performance. This can been seen on Fig. 9. For every learning session during which the PL units responding during the response-reward interval could be monitored in parallel with LC activity, as exemplified in Fig. 9, LC acquisition of a differential response to the new CS+ always preceded behavioural discrimination and the reappearance of PL response-reward interval activity.

Discussion LC activity The striking homogeneity in spontaneous and evoked activity in LC neurons both within and across trials reaffirms what has previously been reported for rats (Sara & Segal, 1991; Valentino et al., 1991; Weiss et al., 1994; Vankov et al., 1995; Shinba et al., 2000) and primates (Aston-Jones et al., 1997, 1999). Virtually all cells showed the same pattern of responses to within-trial events and the same plasticity of responding with changing stimulus-reward contingencies across trials. The fact that these cells fire en masse in a similar mode (tonic or phasic responses) during well-defined aspects of the task, is strong evidence for noradrenergic modulation of the responsivity and

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

Locus coeruleus-frontal cortex interplay 799

Fig. 8. Plasticity of LC response to CS+ during reversal. Raster display for a LC SU activity triggered on vanilla (left) and strawberry (right) onset during a reversal session. Before reversal (top), this LC neuron (mean baseline firing rate ¼ 0.5 ± 0.2 Hz) shows a clear excitatory response to vanilla (CS+, 300% increase from baseline, Wilcoxson: z ¼ )2.7, P ¼ 0.008) and no response to strawberry (CS–). Arrow head indicates mean behavioural response occurrence. At the beginning of reversal, the animal shows an extinction and does not react to any odour. Note the rapid decrease in LC response to vanilla. The first time the animal shows a Go response to strawberry odour and gets the reward, LC neurons show a clear excitation at the time of reward delivery (arrow). The raster presentation on the right shows LC activity triggered on reward delivery during the reversal (dotted line, average nose-poke). LC neurons clearly respond first to reward (700% increase in firing rate, for the first ten trials, Wilcoxson: z ¼ )2.5, P ¼ 0.01) and not to CS+ (z ¼ 0.8, P ¼ 0.4) before this response shifts to the CS+ odour. The reversal of LC response to odours is faster than the behavioural reversal, the first signs of which only appear after 45 trials (see vertical bar on the left indicating period of behavioural discrimination). The response to odours before and after reversal was compared: they significantly reversed (Friedman anova v2(3) ¼ 7.7, P ¼ 0.05) and a posthoc Newman–Keuls analogue test revealed that the response to strawberry at the end of reversal (300% increase from baseline) is not different from the response to vanilla before reversal.

plasticity of forebrain networks mediating cognitive processes underlying behavioural output. The short duration, phasic response to the orienting signal, the light, by virtually all LC cells is similar to that reported in previous experiments in rats and primates, although the latency seen here (155 ms) is considerably longer than that previously reported (Aston-Jones & Bloom, 1981). The LC response to light endured over trials, with no sign of habituation, as long as it signalled the arrival of a discriminative CS, whether or not the odour–reward association was clearly established. On the other hand, this response disappeared during extinction when the animal stopped orienting to the light (and disengaged from the task). Moreover, it can be dissociated from the response to CS+, which disappeared earlier during extinction (within five trials). Therefore, this phasic LC activation seems to be related to the orientation of attention elicited by the light. The interval between the light onset, announcing the beginning of the trial, and the CS presentation, varied between 1.5 and 2.5 s; even so, the data revealed an unexpected significant and reliable decrease in LC firing rate in the hundreds of milliseconds immediately preceding the presentation of the CS (Fig. 3). It is tempting to relate this previously unnoted phenomenon to the ‘expectancy’ mode of attention, a behavioural state shown to be dependent upon inhibition of NA activity in the cat (Bouyer et al., 1989; Delagrange et al., 1989).

The LC differential response to olfactory CSs is similar to earlier reports from our laboratory for auditory CSs in rats (Sara & Segal, 1991) and for auditory and visual stimuli in monkeys (Aston-Jones et al., 1997). As in previous experiments, the adaptation of the response in LC neurons to new stimulus-reward contingencies occurred very rapidly, long before any behavioural expression of the extinction or reversal (Sara & Segal, 1991; Aston-Jones et al., 1997). The response latency to the odour CS+ is relatively long (257 ms), but examination of the data reveals that the LC response is more closely associated with the behavioural response of nose-poking the reward port than with the appearance of the CS+ , itself (Fig. 4). This could account for the discrepancy between the present results and previous results from our laboratory, where LC response to CS+ habituated during stable performance. In those experiments, reward delivery followed CS+ independently of the rat’s behaviour, in a strictly Pavlovian manner (Sara & Segal, 1991). Therefore, persistent LC activation seems to be related to the reward directed behavioural response. At the beginning of each reversal, postextinction relearning or IDS session, where the former CS+ no longer predicts reward, LC responds to the reward, itself. This response rapidly shifts to the CS+ that is predictive of the reward, many trials before behavioural expression of learning of the new stimulus-reward contingency. This is reminiscent of the responses of dopamine (DA) neurons during conditioning as reported extensively by Schultz, but there

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

800 S. Bouret and S. J. Sara

Fig. 9. Plasticity of PL and LC responses during IDS. (Left) Eight neurons displaying reward anticipation activity (during the interval between behavioural response and reward delivery) were monitored during IDS. Mean reward-anticipation activity (black line) and baseline activity (grey line) of these eight cells were calculated during four blocks of 20 trials: before IDS, right before criteria, first block at criteria and second block after criteria. A non-parametric Friedman anova performed over the four blocks confirmed that the reward anticipation activity changed during IDS (v2(7) ¼ 34.5, P < 0.0001). Mean reward-anticipation activity is significantly different from baseline before IDS and when criterion has been reached (posthoc Newman–Keuls analogue, P < 0.05). The response was not significant at the beginning of IDS, before the animal expressed a behavioural discrimination. (Right) PSTH (50 ms bin) and raster display triggered on reward delivery for the activity of a LC unit (1.4 ± 0.2 Hz baseline activity, left) simultaneously recorded with a PL unit (0.35 ± 0.1 Hz baseline activity, right). Solid line and broken line indicate average odour onset and behavioural response, respectively). Top PSTHs are constructed from an asymptotic performance session. The LC excitatory response to CS+ (71% increase in firing rate), related to the behavioural response, is clearly visible, starting approximately one second before the reward delivery, i.e. before the response-reward interval. The PL unit shows a clear excitatory response during the interval preceding reward (400% increase from baseline firing rate), when the animal is in the reward port waiting for the reward. The bottom part of the figure shows PSTHs constructed from the IDS session immediately following, when novel odours are introduced. After the animal obtained the reward, which elicited a clear LC excitation for a few trials (arrowhead) (150% increase from baseline firing rate for the first ten trials), it showed a short phase (ten trials) of systematic Go responses to both odours. During that period, LC neurons rapidly (within one trial after first reward) develop a response to CS+, while PL response only re-appears, after the animal reached stable correct performance (bold vertical bar on the right). LC response to CS+ clearly precedes PL response-reward interval activity not only within but also between trials during learning.

are subtle differences. DA neurons provide a prediction error signal, by exciting to unpredicted reward signals and inhibiting to omission of expected reward (Schultz, 2001, 2002). In the present experiments, LC neurons do not provide such a prediction error signal, for they do not respond to reward omission (first extinction trials). The present results clearly establish a role for the LC noradrenergic system in processing information concerning reward expectancy and the incentive value of the CS, extending previous findings emphasizing a role in vigilance (Aston-Jones et al., 1994) and behavioural flexibility (Sara & Segal, 1991; Vankov et al., 1995; Aston-Jones et al., 1997). PL activity Two phenomena were particularly notable in PL responses: first, a tonic inhibition over the entire trial and second, reward related activity (before or after reward delivery). Response to the preparatory light stimulus, seen in 35% of PL units, is almost exclusively inhibitory, with the inhibition generally lasting until the CS offset. These data strikingly parallel a recent fMRI study where a homologous human prefrontal region (Subgenual PFC, BA 25) was deactivated in a working memory task,

when the difficulty of the task was increased along with the amount of reward at stake. The authors suggest that this ‘emotional’ cortical region is inhibited when the task demands a high level of attention, allowing more cognitive regions to predominate (Pochon et al., 2002). Interestingly, Schoenbaum & Eichenbaum (1995a,b) found cells with tonic inhibitory responses in rat orbital frontal cortex (OfCx) in similar proportion to that of the present experiments, using a behavioural protocol similar to ours. Taken together, these data in rats and humans suggest that neural activity in several limbic areas is tonically inhibited when some attentional processes are engaged. In agreement with other studies (Takenouchi et al., 1999; Gill et al., 2000; Pratt & Mizumori, 2001), we found a population of PL cells activated when the animal was expecting the reward, during the delay between the behavioural response and the actual reward delivery. Another population, partially overlapping with the former, was engaged at reward delivery. This reward and reward expectancy activity might be compared to data obtained in cingulate cortex (Shidara & Richmond, 2002; Koyama et al., 2001; Akkal et al., 2002), or OfCx (Tremblay & Schultz, 2000) in the primate. Moreover, the proportion of cells showing reward expectancy activity is strikingly similar to that described by Schoenbaum et al. (1998) in OfCx and BLA in rat (25%). In line with their report, we found that rewardexpectancy activity in PL neurons shows a relatively slow rate of

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

Locus coeruleus-frontal cortex interplay 801 plasticity during learning. When rats were confronted with a new odour-reward contingency, the reward-expectancy activity disappeared during the early phases of learning and only reappeared after animals showed consistent correct responses. These results are similar to those of a previous report showing the emergence of a tonic reward anticipation response in mFCx only after rats had reached a learning criterion (Mulder et al., 2003). The present data extend these observations by dissociating a tonic, attention-related inhibition, initiated by light onset, from a specific neuronal response occurring during the interval between the behavioural response and the reward. The common relevant finding of these studies is that reward-anticipation activity in PL neurons occurs only after the animal has behaviourally expressed learning of the response-reward contingency. Another population of cells, partially overlapping with the one engaged during reward anticipation, was engaged at reward delivery, thus providing a complementary signal for trial outcome. The task-related neuronal activity shown in the present study, along with that reported by Mulder et al. (2003) provides support to the emerging view of the role of rat mPFC in goal-directed behaviour, largely based on lesion studies. (Balleine & Dickinson, 1998; Birrell & Brown, 2000; De Bruin et al., 2000; Delatour & Gisquet-Verrier, 2000; Killcross & Coutureau, 2003). LC-PL interplay LC activation within the trial occurs reliably at the moment of cognitive shifts required by the protocol: (i) at light onset (when the animal shows an orienting response, interrupting other behaviours to move to the odour port) and (ii) after CS+ (right before the rewarddirected behavioural response, as clearly seen in Fig. 3). Importantly, mFCx responses always follow LC activation, although LC does not systematically drive mFCx responses (several cells show inhibition to the light and excitation during reward anticipation). The neuronal responses observed in PL have also been observed in OfCx and BLA, as well as in piriform cortex (Schoenbaum & Eichenbaum, 1995a,b; Schoenbaum et al., 1998). Furthermore, late modification of EEG in the olfactory bulb in relation to CS expectation or delivery has been observed in rats engaged in a similar task (Ravel et al., 2003). Thus, it appears that many regions of the brain show a systematic modification of activity at these transition phases, and all of these regions receive massive LC projections. Furthermore, a recent fMRI study in humans emphasized the impact of NA on functional interaction between multiple areas, this influence being task and state dependent (Coull et al., 1999). Thus, LC activation at transitional phases within the trial might act as a ‘reset signal’, promoting functional modifications of ensemble activity throughout the underlying networks. This resetting may be an essential condition for the behavioural shift observed after light onset (orientation) or CS+ (reward-directed behaviour). In conclusion, we show that LC neurons respond reliably and homogeneously to both attention orienting (light) and reward-associated signals (CS+) within the trial. Furthermore, during learning, LC neurons acquire the new odour-reward association before the animal expresses it behaviourally, while PL neuronal adaptation is only observed after the behavioural learning. Strong evidence from experiments using a variety of approaches (Schoenbaum et al., 1998; Tronel & Sara, 2002; Cardinal et al., 2003) suggests that PL is part of a functional network, including amygdaloid nuclei and OfCx, involved in various aspects of odour discrimination learning. It is very likely that LC activation to both the orienting signal and the reward-associated signal modulates neuronal processing during learning, within each area, and between functional

neuronal ensembles distributed in the ubiquitous target regions of LC neurons. Thereby behavioural adaptation to changing environmental imperatives would be facilitated.

Acknowledgements We should like to thank Y. Moricard for excellent histology as well as for help in preparation of the manuscript. We also thank E. Kublik for helpful advice and J. C. Lecas for comments on data analysis and interpretation. This work was supported by CNRS UMR 7102. S.B. was supported by a predoctoral grant from the French Ministe`re de l’Education Nationale, Recherche et Technologie and by a grant from the Fondation pour la Recherche Medicale.

Abbreviations IDS, intra-dimensional shift; LC, locus coeruleus; mFCx, medial frontal cortex; MU, multiple unit recording; NA, noradreniline; OfCx, orbital frontal cortex; PFC, prefrontal cortex; PL, prelimbic cortex; PSTHs, post stimulus time histograms; SU, single units.

References Akkal, D., Bioulac, B., Audin, J. & Burbaud, P. (2002) Comparison of neuronal activity in the rostral supplementary and cingulate motor areas during a task with cognitive and motor demands. Eur. J. Neurosci., 15, 887–904. Arnsten, A.F. (1997) Catecholamine regulation of the prefrontal cortex. J. Psychopharmacol., 11, 151–162. Aston-Jones, G. & Bloom, F.E. (1981) Norepinephrine-containing locus coeruleus neurons in behaving rats exhibit pronounced responses to nonnoxious environmental stimuli. J. Neurosci., 1, 887–900. Aston-Jones, G., Rajkowski, J. & Cohen, J. (1999) Role of locus coeruleus in attention and behavioral flexibility. Biol. Psychiatry, 46, 1309–1320. Aston-Jones, G., Rajkowski, J. & Kubiak, P. (1997) Conditioned responses of monkey locus coeruleus neurons anticipate acquisition of discriminative behavior in a vigilance task. Neuroscience, 80, 697–715. Aston-Jones, G., Rajkowski, J., Kubiak, P. & Alexinsky, T. (1994) Locus coeruleus neurons in monkey are selectively activated by attended cues in a vigilance task. J. Neurosci., 14, 4467–4480. Balleine, B.W. & Dickinson, A. (1998) Goal directed instrumental action: contingency and incentive learning and their cortical substrates. Neuropharmacology, 37, 407–419. Berridge, C.W. & Waterhouse, B.D. (2003) The locus coeruleus-noradrenergic system: modulation of behavioral state and state-dependent cognitive processes. Brain Res. Brain Res. Rev., 42, 33–84. Birrell, J.M. & Brown, V.J. (2000) Medial frontal cortex mediates perceptual attentional set shifting in the rat. J. Neurosci., 20, 4320–4324. Bouret, S. & Sara, S.J. (2002) Locus coeruleus activation modulates firing rate and temporal organization of odour-induced single-cell responses in rat piriform cortex. Eur. J. Neurosci., 16, 2371–2382. Bouyer, J.J., Tilquin, C., Delagrange, P., Rougeul, A. & Buser, P. (1989) Regulation pathways of expectancy behavior in the cat: histologic study. C. R. Acad. Sci., III, 309, 637–646. Cardinal, R.N., Parkinson, J.A., Marbini, H.D., Toner, A.J., Bussey, T.J., Robbins, T.W. & Everitt, B.J. (2003) Role of the anterior cingulate cortex in the control over behavior by Pavlovian conditioned stimuli in rats. Behav. Neurosci., 117, 566–587. Coull, J.T., Buchel, C., Friston, K.J. & Frith, C.D. (1999) Noradrenergically mediated plasticity in a human attentional neuronal network. Neuroimage, 10, 705–715. De Bruin, J.P., Feenstra, M.G., Broersen, L.M., Van Leeuwen, M., Arens, C., De Vries, S. & Joosten, R.N. (2000) Role of the prefrontal cortex of the rat in learning and decision making: effects of transient inactivation. Prog. Brain Res., 126, 103–113. Delagrange, P., Tadjer, D., Bouyer, J.J., Rougeul, A. & Conrath, M. (1989) Effect of DSP4, a neurotoxic agent, on attentive behaviour and related electrocortical activity in cat. Behav. Brain Res., 33, 33–43. Delatour, B. & Gisquet-Verrier, P. (2000) Functional role of rat prelimbicinfralimbic cortices in spatial memory: evidence for their involvement in attention and behavioural flexibility. Behav. Brain Res., 109, 113–128. Florin-Lechner, S.M., Druhan, J.P., Aston-Jones, G. & Valentino, R.J. (1996) Enhanced norepinephrine release in prefrontal cortex with burst stimulation of the locus coeruleus. Brain Res., 742, 89–97.

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

802 S. Bouret and S. J. Sara Gill, T.M., Sarter, M. & Givens, B. (2000) Sustained visual attention performance-associated prefrontal neuronal activity: evidence for cholinergic modulation. J. Neurosci., 20, 4745–4757. Harley, C.W., Lacaille, J.C. & Galway, M. (1983) Hypothalamic afferents to the dorsal dentate gyrus contain acetylcholinesterase. Brain Res., 270, 335–339. Killcross, S. & Coutureau, E. (2003) Coordination of actions and habits in the medial prefrontal cortex of rats. Cereb. Cortex, 13, 400–408. Koyama, T., Kato, K., Tanaka, Y.Z. & Mikami, A. (2001) Anterior cingulate activity during pain-avoidance and reward tasks in monkeys. Neurosci. Res., 39, 421–430. Lecas, J.C. (2004) Locus coeruleus activation shortens synaptic drive, decreases latency and improves spike timing precision in sensorimotor cortex. Implication for neuronal integration. Eur. J. Neurosci., 19, 2519–2530. Mantz, J., Milla, C., Glowinski, J. & Thierry, A.M. (1988) Differential effects of ascending neurons containing dopamine and noradrenaline in the control of spontaneous activity and of evoked responses in the rat prefrontal cortex. Neuroscience, 27, 517–526. ¨ rgu¨t, O. & Pennartz, C.M.A. (2003) LearningMulder, A.B., Nordquist, B., O related changes in response patterns of prefrontal neurons during instrumental conditioning. Behav. Brain Res., 146, 77–88. Passetti, F., Chudasama, Y. & Robbins, T.W. (2002) The frontal cortex of the rat and visual attentional performance: dissociable functions of distinct medial prefrontal subregions. Cereb. Cortex, 12, 1254–1268. Pochon, J.B., Levy, R., Fossati, P., Lehericy, S., Poline, J.B., Pillon, B., Le Bihan, D. & Dubois, B. (2002) The neural system that bridges reward and cognition in humans: an fMRI study. Proc. Natl Acad. Sci. USA, 99, 5669–5674. Pratt, W.E. & Mizumori, S.J. (2001) Neurons in rat medial prefrontal cortex show anticipatory rate changes to predictable differential rewards in a spatial memory task. Behav. Brain Res., 123, 165–183. Ravel, N., Chabaud, P., Martin, C., Gaveau, V., Hugues, E., Tallon-Baudry, C., Bertrand, O. & Gervais, R. (2003) Olfactory learning modifies the expression of odour-induced oscillatory responses in the gamma (60–90 Hz) and beta (15–40 Hz) bands in the rat olfactory bulb. Eur. J. Neurosci., 17, 350–358. Sara, S.J. & Segal, M. (1991) Plasticity of sensory responses of locus coeruleus neurons in the behaving rat: implications for cognition. Prog. Brain Res., 88, 571–585. Sawaguchi, T. & Kikuchi, Y. (1998) Noradrenergic effects on activity of prefrontal cortical neurons in behaving monkeys. Adv. Pharmacol., 42, 759–763. Schoenbaum, G., Chiba, A.A. & Gallagher, M. (1998) Orbitofrontal cortex and basolateral amygdala encode expected outcomes during learning. Nature Neurosci., 1, 155–159.

Schoenbaum, G. & Eichenbaum, H. (1995a) Information coding in the rodent prefrontal cortex. I. Single- neuron activity in orbitofrontal cortex compared with that in pyriform cortex. J. Neurophysiol., 74, 733–750. Schoenbaum, G. & Eichenbaum, H. (1995b) Information coding in the rodent prefrontal cortex. II. Ensemble activity in orbitofrontal cortex. J. Neurophysiol., 74, 751–762. Schultz, W. (2001) Reward signaling by dopamine neurons. Neuroscientist, 7, 293–302. Schultz, W. (2002) Getting formal with dopamine and reward. Neuron, 36, 241–263. Shidara, M. & Richmond, B.J. (2002) Anterior cingulate: single neuronal signals related to degree of reward expectancy. Science, 296, 1709– 1711. Shinba, T., Briois, L. & Sara, S.J. (2000) Spontaneous and auditory-evoked activity of medial agranular cortex as a function of arousal state in the freely moving rat: interaction with locus coeruleus activity. Brain Res., 887, 293– 300. Takenouchi, K., Nishijo, H., Uwano, T., Tamura, R., Takigawa, M. & Ono, T. (1999) Emotional and behavioral correlates of the anterior cingulate cortex during associative learning in rats. Neuroscience, 93, 1271–1287. Tremblay, L. & Schultz, W. (2000) Modifications of reward expectation-related neuronal activity during learning in primate orbitofrontal cortex. J. Neurophysiol., 83, 1877–1885. Tronel, S. & Sara, S.J. (2002) Mapping of olfactory memory circuits: regionspecific c-fos activation after odor-reward associative learning or after its retrieval. Learn. Mem., 9, 105–111. Valentino, R.J., Page, M.E. & Curtis, A.L. (1991) Activation of noradrenergic locus coeruleus neurons by hemodynamic stress is due to local release of corticotropin-releasing factor. Brain Res., 555, 25–34. Vankov, A., Herve´-Minvielle, A. & Sara, S.J. (1995) Response to novelty and its rapid habituation in locus coeruleus neurons of the freely exploring rat. Eur. J. Neurosci., 7, 1180–1187. Walling, S.G. & Harley, C.W. (2004) Locus ceruleus activation initiates delayed synaptic potentiation of perforant path input to the dentate gyrus in awake rats: a novel beta-adrene. J. Neurosci., 24, 598–604. Waterhouse, B.D. & Woodward, D. (1980) Interaction of norepinephrine with cerebrocortical activity evoked by stimulation of somatosensory afferent pathways in the rat. Exp. Neurol., 67, 11–34. Weiss, G.K., Ratner, A., Voltura, A., Savage, D., Lucero, K. & Castillo, N. (1994) The effect of two different types of stress on locus coeruleus alpha-2 receptor binding. Brain Res. Bull., 33, 219–221.

ª 2004 Federation of European Neuroscience Societies, European Journal of Neuroscience, 20, 791–802

Reward expectation, orientation of attention and locus ...

Go responses to the CS– odour (commission errors) were signalled by the light being ... >90% correct response for both CS+ and CS– odours, before electrodes were ...... Schoenbaum, G., Chiba, A.A. & Gallagher, M. (1998) Orbitofrontal cortex and basolateral ... signals related to degree of reward expectancy. Science, 296 ...

500KB Sizes 1 Downloads 204 Views

Recommend Documents

Reward expectation, orientation of attention and locus ...
throughout training and recording sessions and rats were maintained at approximately .... software. Further spike sorting was made off-line. If the spikes were.

Scope of attention, control of attention, and intelligence ...
that variance was shared between scope and control, and the rest was unique to one or the other. Scope and control of ... ward an object appearing on the screen, an “antisaccade” movement ...... service of memory that was poorer in children.

Locus iste
Locus iste. Music by: Anton Bruckner (1824-1896). WAB 23 original key: C. Allegro moderate ._ .. ._ _= I. _ — D - .__ D : DDD n= D f:- D . . .= _ =. _ 5 . E _ _.

Rhetoric and Practice of Strategic Reward Management.pdf ...
Rhetoric and Practice. of. Strategic. Reward Management. Page 3 of 510. Rhetoric and Practice of Strategic Reward Management.pdf. Rhetoric and Practice of ...

Rhetoric and Practice of Strategic Reward Management.pdf ...
Rosario Longo. Rhetoric and Practice. of. Strategic. Reward Management. Executive Pay. Rewarding Teams. Profit-Sharing. Cafeteria Benefits. Execution. Total Reward. Reward Risk. Flexible Benefits. Voluntary Benefits. Pension Plans. Gain-sharing. Pay

CONDITIONAL MEASURES AND CONDITIONAL EXPECTATION ...
Abstract. The purpose of this paper is to give a clean formulation and proof of Rohlin's Disintegration. Theorem (Rohlin '52). Another (possible) proof can be ...

A multi-locus phylogeny of Nectogalini shrews and ...
c Department of Ecology and Evolutionary Biology, Yale University, New Haven, CT, USA d Laboratory of Wildlife Ecology, ... Available online 2 April 2010. Keywords: ...... tree space is potential mis-alignment of the rRNA of mtDNA. How-.

dependence, locus of control, parental bonding, and ... - Oxford Journals
AND PERSONALITY DISORDERS: A STUDY IN ALCOHOLICS. AND CONTROLS .... found in the level of education (P < 0.05) and the ... Professional school (%). 12.1. 14.7 .... Schuckit, M. A., Klein, J. B. S., Twitchell, G. B. S. and. Smith T.

Search, Effort, and Locus of Control Andrew McGee ...
University of Arkansas. February 2016 ... Address: Department of Economics, University of Arkansas, 402 Business. Building, Fayetteville ..... beliefs, other characteristics, and a linear time trend.7 In all of the tables discussed in this section, w

Search, Effort, and Locus of Control Andrew McGee ...
imaginativeness or intellect (Goldberg, 1992), the Rosenberg Self-Esteem Scale (Rosenberg, 1965), and a ..... services), the number of job applications submitted (Caliendo et al. 2015), the ..... Princeton: Princeton University Press. Rotter ...

DERIVED EQUIVALENCES OF K3 SURFACES AND ORIENTATION ...
As the image of a Kähler class is again a Kähler class, one actually has f∗(CX) = CX. In other words, f∗ respects the connected components of the set of (1,1)-classes α with (α, α) > 0. If one wants to avoid the existence of Kähler structur

Expectation Formation Rules and the Core of Partition ...
Sep 20, 2013 - Definitions of the core of partition function games proposed in the litera- ..... symmetry considerations for the optimistic, pessimistic, and max ...

Discussion of Liquidity Traps and Expectation ...
Sep 20, 2013 - Backup slides. 24 / 25. Page 25. Temporary equilibrium equations. New Keynesian Phillips curve: Q(πt) = (πt − 1)πt. Q(πt) = ν γα y. 1+ǫ α t. −.

Repetition Suppression and Expectation Suppression ...
Sep 26, 2012 - data analysis that may arise from having an unequal number of trials per condition ... were performed using SPSS Statistics v16. (IBM).

Repetition Suppression and Expectation Suppression Are Dissociable ...
Sep 26, 2012 - we observed ES (but no RS) during an intermediate (100–200 .... domain. More specifically, for the early time window, we chose to focus on the P50 ..... Friston K (2009) The free-energy principle: a rough guide to the brain?

Expectation, Disappointment, and Exit: Reference ... - Wharton Marketing
Dec 29, 2016 - way we make policy, manage firms, and design markets. ... and endogenously within the context of the application.1 A practical theory of reference- .... Importantly, the BIN option may disappear in the course of the auction.

Attention and Celebration.pdf
Whoops! There was a problem loading more pages. Whoops! There was a problem previewing this document. Retrying... Download. Connect more apps.

Intersection of Reward and Memory in Monkey Rhinal Cortex
May 16, 2012 - B,C, Schematic illustration of the behavioral tasks. In the visually cued ... square) 50 ms after bar release and reward de- livery 200 – 400 .... ment, or within-block trial number had a significant effect on perfor- mance. Because 

Cognitive control of attention and action: Issues and ...
Accepted: 18 March 2002 / Published online: 14 August 2002. © Springer-Verlag 2002 ... able to configure and re-configure their cognitive system in a way that .... the recruitment and/or intervention of control and organization of basic cognitive ..

Attention and Celebration.pdf
Page 2 of 2. Pretend to squirt cleaner on window with sound three times. Then. circular wipe motions say FANTASTIC! Stomp, stomp, clap. Stomp, stomp, clap.

The social brain and reward
provided a potential neurophysiological explanation for computational models describing how individuals learn .... the popular opinion agrees rather than disagrees with our own evaluation of an object 44, 45, and is related to our .... they exhibit g

Information and attention
Dec 6, 2011 - a speaker is taken to provide the information that the actual world is one ...... and the question in (ii-b) are quantifying operators, and the domain.

Testosterone exposure, dopaminergic reward, and sensation ...
Testosterone exposure, dopaminergic reward, and sensation seeking in young men.pdf. Testosterone exposure, dopaminergic reward, and sensation seeking in ...