Zum Hauptinhalt springen

Effects of intra-accumbal administration of dopamine and ionotropic glutamate receptor drugs on delay discounting performance in rats.

Yates, JR ; Bardo, MT
In: Behavioral neuroscience, Jg. 131 (2017-10-01), Heft 5, S. 392
Online academicJournal

Effects of Intra-Accumbal Administration of Dopamine and Ionotropic Glutamate Receptor Drugs on Delay Discounting Performance in Rats By: Justin R. Yates
Department of Psychology, University of Kentucky
Michael T. Bardo
Department of Psychology, University of Kentucky;

Acknowledgement: Michael T. Bardo,
The data presented in the current manuscript were collected as part of Justin R. Yates’ doctoral dissertation. The research was funded by NIH Grants P50 DA05312 and T32 DA007304.

Understanding the neurobiological basis of impulsive choice is important for designing effective treatment options for those with disorders characterized by increased impulsivity, such as attention-deficit/hyperactivity disorder (ADHD) and substance use disorders. Dopamine (DA) has received considerable attention because drugs that are efficacious in treating ADHD exert their effects by increasing DA, as well as norepinephrine, levels (see Bidwell, McClernon, & Kollins, 2011 for a review). Preclinical models have shown that systemic administration of DA D1 receptor antagonists increase impulsive choice (Broos, Diergaarde, Schoffelmeer, Pattij, & De Vries, 2012; Koffarnus, Newman, Grundt, Rice, & Woods, 2011; van Gaalen, van Koten, Schoffelmeer, & Vanderschuren, 2006; but see Wade, de Wit, & Richards, 2000), although results following antagonism of D2 receptors have been mixed. A couple of studies reported increases in impulsive choice following antagonism of D2 receptors (Denk et al., 2005; Wade et al., 2000), but others found no alterations in discounting (Evenden & Ryan, 1996; Koffarnus et al., 2011; van Gaalen et al., 2006).

In addition to DA, recent evidence has implicated glutamate (Glu) dysfunction in impulse-control disorders, including ADHD (Jensen et al., 2009; Miller, Pomerleau, Huettl, Gerhardt, & Glaser, 2014; Perlov et al., 2007) and substance use disorders (Ben-Shahar et al., 2012; Griffin, Haun, Hazelbaker, Ramachandra, & Becker, 2014). Directly related to impulsive choice, blocking N-methyl-D-aspartate (NMDA) receptors with the uncompetitive antagonists memantine and ketamine increases delay discounting (Cottone et al., 2013; Floresco, Tse, & Ghods-Sharifi, 2008; but see Yates, Gunkel, Rogers, Hughes, & Prior, 2017), whereas the uncompetitive antagonist MK-801 decreases discounting (Higgins et al., 2016; Yates, Batten, Bardo, & Beckmann, 2015; but see Yates, Gunkel et al., 2017). Similar to MK-801, antagonists at NR2B-containing NMDA receptors decrease impulsive choice (Higgins et al., 2016; but see Yates, Gunkel et al., 2017).

Although there is evidence to support a role for DA and Glu in impulsive choice using systemic injections, few studies have examined the neuroanatomical regions that control the effects of DA on impulsive decision making. To date, research has shown that direct infusions of D1 receptor antagonists (Pardey, Kumar, Goodchild, & Cornish, 2013; Loos et al., 2010) and D2 receptor antagonists (Pardey et al., 2013; Yates et al., 2014) into medial prefrontal cortex (mPFC) increase impulsive choice. Although blocking D2 receptors in orbitofrontal cortex (OFC) does not typically alter impulsive choice (Pardey et al., 2013; Yates et al., 2014), Zeeb, Floresco, and Winstanley, 2010 found that a D2 receptor antagonist increased impulsive choice, but only when the delay to the delivery of the large reinforcer was signaled, suggesting that environmental cues can moderate the effects of DA receptor ligands on impulsivity. To our knowledge, the effects of direct infusions of Glu receptor drugs on impulsive choice have not been examined.

One important neural mediator of impulsive choice, specifically delay discounting, is the nucleus accumbens core (NAcc). Lesions to NAcc increase preference for a small, immediate reinforcer relative to a larger, delayed reinforcer (Bezzina et al., 2007; Cardinal, Pennicott, Sugathapala, Robbins, & Everitt, 2001; da Costa Araújo et al., 2009; Galtress & Kirkpatrick, 2010; Pothuizen, Jongen-Rêlo, Feldon, & Yee, 2005; Valencia-Torres et al., 2012; but see Winstanley, Theobald, Dalley, & Robbins, 2005), although inactivation with GABA receptor agonists has been reported to either increase (Feja, Hayn, & Koch, 2014) or decrease (Moschak & Mitchell, 2014) impulsive choice. There is some evidence that NAcc DA plays an important role in impulsive choice, as overexpression of the DA transporter (DAT) leads to increased impulsivity (Adriani et al., 2009). Additionally, Winstanley et al. (2005) argue that DA and serotonin (5-HT) interactions within NAcc contribute to impulsive choice, as the increased impulsivity following 8-OH-DPAT (5-HT1A agonist) administration is not observed in DA-depleted rats. Despite the evidence for a role of NAcc in discounting, the specific contribution of DA or Glu receptors within this region is largely unknown.

The goal of the present study was to determine the contribution of NAcc DA and Glu receptors to delay discounting. Because sensitivity to delayed reinforcement and sensitivity to reinforcer magnitude independently influence discounting of a reinforcer (Ho, Mobini, Chiang, Bradshaw, & Szabadi, 1999), we applied quantitative analyses (e.g., exponential discounting function) to determine if DA receptors and Glu receptors within NAcc differentially alter these parameters. Quantitative analyses have been used previously to determine the effects of excitotoxic lesions (Bezzina et al., 2007, 2008, 2009; da Costa Araújo et al., 2009; Kheramin et al., 2002; Valencia-Torres et al., 2012) and pharmacological manipulations (Yates et al., 2015; Yates, Gunkel et al., 2017, Yates, Rogers et al., 2017) on delay discounting. By applying quantitative analyses in the current study, we sought to determine the precise mechanisms by which NAcc DA or Glu receptors mediate discounting.

Method
Subjects

A total of 24 male, individually housed Sprague–Dawley rats (Harlan Industries; Indianapolis, IN) were used in the experiments. Rats weighed approximately 250–275g (approximately postnatal Day 60) upon arrival to the laboratory. Rats were acclimated to a colony room held at a constant temperature and were handled for 5 days upon arrival. Light and dark phases were on a 12-hr light–dark cycle, and each experiment occurred during the light phase. Rats were food restricted (approximately 80% of free feed body weight) 3 days before the beginning of behavioral training, and rats remained on food restriction during the remainder of the study, unless otherwise noted. Rats were cared for in accordance with the Guide for the Care and Use of Laboratory Animals (National Research Council, 2011), and procedures were approved by the Institutional Animal Care and Use Committee at the University of Kentucky.

Drugs

The following were purchased from Sigma Aldrich (St. Louis, MO): (±)-SKF-38393 hydrochloride, (±)-SCH-23390 hydrochloride, (-)-quinpirole hydrochloride, S-(-)-eticlopride hydrochloride, (+)-MK-801 hydrogen maleate, D(-)-2-amino-5-phosphonopentanoic acid (AP-5), and 6-cyano-7-nitroquinoxaline-2,3-dione disodium salt hydrate (CNQX). Ifenprodil hemitartrate was purchased from Tocris Bioscience (Ellisville, MO). Each drug was prepared in sterile 0.9% NaCl (saline), except for ifenprodil, which was prepared in sterile water. Concentrations were calculated based on salt weight.

Behavioral Apparatus

Operant conditioning chambers (28 × 21 × 21 cm; ENV-008; MED Associates, St. Albans, VT) located inside sound-attenuating chambers (ENV-018M; MED Associates) were used. The front and back walls of the experimental chambers were made of aluminum, while the side walls were made of Plexiglas. There was a recessed food tray (5 × 4.2 cm) located 2 cm above the floor in the bottom-center of the front wall. An infrared photobeam was used to record head entries into the food tray. A 28-V white cue light was located 6 cm above each response lever. A white house light was mounted in the center of the back wall of the chamber. All responses and scheduled consequences were recorded and controlled by a computer interface. A computer controlled the experimental session using Med-IV software.

Procedure

Rats were given 2 days of magazine training, in which sucrose-based 45 mg pellets (F0021 dustless precision pellet, Bio-Serve, Frenchtown, NJ) were noncontingently delivered into the food tray. These sessions were used to habituate rats to the operant chamber. Following magazine training, rats were given lever press training. Each session began with illumination of the house light. A head entry into the food hopper resulted in presentation of one lever. Levers were presented semirandomly, with no more than two consecutive presentations of the same lever. A response on either lever resulted in delivery of one sucrose pellet. Pellets were also delivered noncontingently on a random time 100-s schedule of reinforcement. Following a response on either lever, the house light was extinguished, and the lever was retracted for 5 s. After 5 s, the house light was illuminated. Each session lasted 30 min or following completion of 40 trials, whichever occurred first.

After three sessions, rats received reward magnitude discrimination training, which consisted of 40 trials each day. Each trial lasted 40 s and began with illumination of the house light. A head entry into the food hopper extended one of the levers (semirandomly presented, with no more than two consecutive presentations of the same lever). A response on one lever resulted in immediate delivery of one pellet, whereas a response on the other lever resulted in immediate delivery of four pellets (the lever associated with the large reward magnitude was counterbalanced across rats). Following a response, the house light was extinguished, and the lever was retracted for the remainder of the trial. If a response was not made within 10 s, the trial was scored as an omission, and the house light was extinguished for the remainder of the trial. After 7 days of reward magnitude discrimination training, rats were trained in delay discounting sessions.

Delay discounting sessions consisted of five blocks of nine trials, and each trial lasted 60 s. The first four trials in a block were forced-choice trials, in which only one lever was semirandomly presented (no more than two consecutive presentations of the same lever). The last five trials were free-choice trials, in which both levers were extended. As in reward magnitude discrimination training, a response on one lever always resulted in immediate delivery of one food pellet. A response on the other lever resulted in delivery of four pellets; however, the delay to the delivery of the large magnitude reward increased across blocks of trials (0, 5, 10, 20, 50 s). Following a response on either lever, the house light was extinguished, and the lever was retracted for the remainder of the trial. If a response was not made within 10 s, the trial was scored as an omission, and the house light was extinguished for the remainder of the trial.

Surgery

After 32 sessions of delay discounting, rats were treated with the nonopioid analgesic carprofen (5 mg/kg, s.c.) 1 day prior to and on the day of surgery. Rats were anesthetized with a mixture of ketamine, xylazine, and acepromazine (75, 7.5, and 0.75 mg/kg, i.p., respectively) and were secured into a stereotaxic frame. Cannulas were implanted bilaterally into NAcc (+1.6 AP, ±1.5 ML, −5.5 DV) at the 10° angle off the midline (Paxinos & Watson, 1998). Following surgery, rats were treated with carprofen for 2 days.

Intracranial Infusions

Rats recovered for 3–5 days and were food restricted before receiving 12 additional training sessions in the delay-discounting task. This additional training was important to ensure that surgery did not alter discounting. For intracranial infusions, rats were gently restrained by the experimenter, and a stainless steel injection cannula (33 gauge; Small Parts, Inc, Miramar, FL) was inserted 2 mm below the tip of the guide cannulas. Each cannula was connected to a 10 μl syringe (Hamilton, Reno, NV) via PE10 tubing (Small Parts, Inc, Miramar, FL). The Hamilton syringes were mounted on an infusion pump (KDS Scientific, Holliston, MA). Half of the rats (n = 12) received direct infusions of SKF 38393 (D1-like agonist; 0.03, 0.1 μg; Loos et al., 2010; Yates et al., 2014), SCH 23390 (D1-like antagonist; 0.3, 1.0 μg; Loos et al., 2010; Yates et al., 2014; Zeeb et al., 2010), quinpirole (D2-like agonist; 0.3, 1.0 μg; Yates et al., 2014), and eticlopride (D2-like antagonist; 0.3, 1.0 μg; Yates et al., 2014; Zeeb et al., 2010); saline was used as the vehicle for each drug. The other half of the rats (n = 12) received direct infusions of MK-801 (uncompetitive NMDA antagonist; 0.3, 1.0 μg; Bakshi & Geyer, 1998; Zhang, Bast, & Feldon, 2000), AP-5 (competitive NMDA antagonist; 0.3, 1.0 μg; Baldwin, Holahan, Sadeghian, & Kelley, 2000; Dar, 2002; Sombers, Beyene, Carelli, & Wightman, 2009), ifenprodil (antagonist at NR2B-containing NMDA receptors; 0.3, 1.0 μg; Parkes & Balleine, 2013; Laurent & Westbrook, 2008), and CNQX (AMPA antagonist; 0.2 and 0.5 μg; Hitchcott & Phillips, 1997; Mesches, Bianchin, & McGaugh, 1996); saline was used as the vehicle for MK-801, AP-5, and CNQX, whereas sterile water was used as the vehicle for ifenprodil. In total, rats received either nine (DA experiment) or 10 (Glu experiment) infusions (see Tables 1 and 2). Each drug was infused over 2 min at a rate of 0.25 μl/min. Injectors were left in place for 1 min following the infusion. Rats were placed into the operant chamber immediately following the infusion. Treatments were randomly administered for each rat, and rats were given 2 days of washout (no drug microinfusion) following each infusion; on these washout days, rats were tested in delay discounting.
bne-131-5-392-tbl1a.gif
bne-131-5-392-tbl2a.gif

Following the last day of infusions, rats were euthanized, and brains were removed and flash-frozen in chromasolv (Sigma, St. Louis, MO) on dry ice and stored at −80 °C until sectioning was completed. Brain sections (40 μm) were sliced to determine the location of guide cannulas. Probe placements were evaluated according to the atlas of Paxinos and Watson (1998). Only data from rats with correct probe placements in NAcc were used in all statistical analyses.

Statistical Analysis

For baseline data (averaged across final four sessions before surgery and averaged across final four sessions before first infusion), two analyses were used to determine if discounting differed across rats assigned to receive DA or Glu infusions, as well as to determine if surgery had an effect on discounting. First, a mixed factor ANOVA was used, with experiment (DA vs. Glu) as a between-subjects factor and surgery (Pre vs. Post) and delay as within-subjects factors. To probe a significant interaction, separate independent-samples t tests (with Bonferroni correction) were used. Because ANOVAs do not specify how sensitivity to reinforcer magnitude and/or sensitivity to delayed reinforcement have been altered, the exponential discounting function was fit to each subject’s data via nonlinear mixed effects modeling (NLME; Pinheiro, Bates, DebRoy, Sarkar, & R Core Team, 2017). The exponential equation was defined as V = AebD, where V is the subjective value of the reinforcer, A is sensitivity to reinforcer magnitude (the intercept of the function), b represents sensitivity to delayed reinforcement (the slope of the function), and D is the delay to the larger magnitude reinforcer. The NLME model defined A and b as free parameters, surgery as a nominal, within-subjects factor, and delay as a continuous, within-subjects factor. A main effect was probed using contrasts in R.

To determine if intracranial infusions significantly altered omissions, separate Friedman tests were conducted for each drug, and Wilcoxon’s signed-ranked tests (with Bonferroni correction) were used to probe significant main effects, when appropriate. To determine if intracranial infusions of each drug altered discounting, two analyses were used. First, two-way repeated measures ANOVAs were used, with delay and drug concentration as within-subjects factors. Dunnet’s post hoc tests were used to probe a main effect of drug concentration. If there was a significant interaction, separate paired-samples t tests (with Bonferroni correction) were used. Second, separate NLME analyses for each drug were conducted, which defined A and b as free parameters, delay as a continuous, within-subjects factor, drug concentration as a nominal, within-subjects factor, and subject as a random factor. Main effects and interactions were probed using contrasts in R.

To determine if A and b parameter estimates changed across each baseline period (averaged across the two sessions between infusions, as well as across the two sessions before the first infusion and the two sessions following the final infusion), linear trend analyses were conducted for each parameter.

Statistical significance was defined as p < .05 in all cases, with the exception of the Wilcoxon’s signed-ranked tests and independent/paired-samples t tests, in which a Bonferroni correction was applied to control for Type I error. For all ANOVA analyses, degrees of freedom were corrected using Greenhouse-Geisser corrections when sphericity was violated. Additionally, partial eta squared (ηp2) was provided as a measure of effect size, with values of 0.01, 0.06, and 0.14 indicating small, medium, and large effect sizes, respectively (Cohen, 1988).

Results

Figure 1 shows probe placements for rats given DA-selective ligands (Figure 1a) and Glu-selective ligands (Figure 1b). Four rats in the Glu experiment and six rats in the DA experiment had probe placements outside of NAcc, and were thus excluded from further analyses.
bne-131-5-392-fig1a.gif

For baseline data, results of the ANOVA revealed a significant main effect of delay, F(2.088, 25.053) = 36.612, p < .001, ηp2 = .753. There were no other significant main effects or interactions, Fs ≤ 1.107, ps ≥ .364, ηp2s ≤ .084. Results of the NLME analysis showed that sensitivity to reinforcer magnitude and sensitivity to delayed reinforcement did not differ across rats assigned to receive DA or Glu infusions, and the analyses showed that surgery did not alter either parameter, Fs ≤ 3.611, ps ≥ .060 (data not shown).

Administration of AP-5 increased omissions, χ2(2) = 8.000, p = .018, although post hoc tests did not reveal a significant difference between vehicle and any concentration of AP-5. It is important to note that each rat had no omissions following vehicle treatment and that one rat had more than one omission following AP-5 administration only. None of the other intracranial infusions significantly altered the number of omissions, χ2s ≤ 5.158, ps ≥ .076 (see Table 3).
bne-131-5-392-tbl3a.gif

Figure 2 shows the raw proportion of responses for the large delayed reinforcer following direct infusions of SKF 38393 (Figure 2a), SCH 23390 (Figure 2b), quinpirole (Figure 2c), and eticlopride (Figure 2d). As expected, the main effect of delay was significant for each analysis, Fs ≥ 14.524, ps < .001, ηp2s ≥ .744. None of the DA-selective ligands significantly altered responding for the large magnitude reinforcer, Fs ≤ 1.970, ps ≥ .190, ηp2s ≤ .283.
bne-131-5-392-fig2a.gif

Figure 3 shows the A and b parameter estimates derived from the exponential model for the DA-selective drugs, and Table 1 shows the percentage change in each parameter relative to vehicle. Although stimulating DA D1 receptors with SKF 38393 did not alter sensitivity to reinforcer magnitude, F(2, 79) = 0.136, p = .873 (Figure 3a), or sensitivity to delayed reinforcement, F(2, 79) = .248, p = .781 (Figure 3a), blocking these receptors with SCH 23390 (0.3 μg) significantly decreased sensitivity to reinforcer magnitude, F(2, 79) = 3.431, p = .037 (Figure 3b), without altering sensitivity to delayed reinforcement, F(2, 79) = 1.586, p = .211 (Figure 3b). Administration of DA D2 receptor ligands did not alter either parameter: quinpirole, F(2, 79) = 0.428, p = .653; F(2, 79) = 0.693, p = .503 (Figure 3c); eticlopride: F(2, 79) = 0.175, p = .840; F(2, 79) = 0.658, p = .521 (Figure 3d).
bne-131-5-392-fig3a.gif

Figure 4 shows the raw proportion of responses for the large delayed reinforcer following direct infusions of MK-801 (Figure 4a), AP-5 (Figure 4b), ifenprodil (Figure 4c), and CNQX (Figure 4d). Similar to the results obtained with the DA-selective ligands, the main effect of delay was significant for each Glu drug tested, Fs ≥ 18.506, ps < .001, ηp2s ≥ .726, but none of the Glu-selective ligands altered responses for the large magnitude reinforcer, Fs ≤ 3.325, ps ≥ .066, ηp2s ≤ .322.
bne-131-5-392-fig4a.gif

Figure 5 shows the A and b parameter estimates derived from the exponential model for the Glu-selective drugs, and Table 2 shows the percentage change in each parameter relative to vehicle. MK-801 did not alter sensitivity to reinforcer magnitude, F(2, 107) = 0.970, p = .383 (Figure 5a), or sensitivity to delayed reinforcement, F(2, 107) = 2.121, p = .125 (Figure 5a). Similarly, AP-5 did not alter either parameter, F(2, 107) = 0.881, p = .417 and F(2, 107) = 0.931, p = .398, respectively (Figure 5b). When examining the effects of ifenprodil on sensitivity to delayed reinforcement, there was a main effect of dose, F(2, 107) = 3.160, p = .046 (Figure 5c), whereas no differences in sensitivity to reinforcer magnitude emerged, F(2, 107) = 0.252, p = .778 (Figure 5c). Although a main effect of ifenprodil was observed for sensitivity to delayed reinforcement, differences between vehicle and ifenprodil (1.0 μg) only trended toward significance (p = .057). The main effect was due primarily to differences between the two ifenprodil concentrations (p = .035). A closer examination of the ifenprodil results revealed that one rat had a b parameter estimate that was approximately 143% larger relative to the average b parameter estimate, resulting in large variation in b parameter estimates following vehicle infusion. A supplementary analysis was conducted after excluding this rat, and results showed a significant effect of ifenprodil on sensitivity to delayed reinforcement, F(2, 93) = 3.337, p = .040); ifenprodil (1.0 μg) significantly decreased b parameter estimates relative to vehicle (p = .039). CNQX did not alter sensitivity to reinforcer magnitude, F(2, 107) = 0.521, p = .595, or impulsive choice, F(2, 107) = .506, p = .604 (Figure 5d).
bne-131-5-392-fig5a.gif

Figure 6 shows A (Figure 6a) and b (Figure 6b) parameter estimates across each baseline session for rats treated with DA drugs and for rats treated with Glu drugs. A parameter estimates did not change across the experiment for either group of rats, DA: F(1, 58) = .467, p = .497; Glu: F(1, 86) = 3.018, p = .086. Similarly, b parameter estimates did not change across the session, DA: F(1, 58) = .401, p = .529; Glu: F(1, 86) = .002, p = .965.
bne-131-5-392-fig6a.gif

Discussion

The goal of the current study was to determine how accumbal DA receptors and ionotropic glutamate receptors, primarily the NMDA receptor subtype, mediate two distinct features of delay discounting: (a) sensitivity to reinforcer magnitude (A parameter); and (b) sensitivity to delayed reinforcement (b parameter). Results showed that SCH-23390 (0.3 μg) decreased sensitivity to a large magnitude reinforcer, whereas ifenprodil (1.0 μg) decreased sensitivity to delayed reinforcement. These results show that DA and Glu receptors within NAcc differentially mediate two dissociable aspects of discounting performance.

DA D1 and D2 receptors are widely expressed in NAcc (Dubois, Savasta, Curet, & Scatton, 1986; Savasta, Dubois, & Scatton, 1986). In the current study, SCH 23390 (0.3 μg) decreased the A parameter without altering the b parameter, thus suggesting a specific role of DA D1 receptors in sensitivity to reinforcer magnitude. This result is consistent with previous work showing that SCH 23390 decreases food reward (Beninger et al., 1987; Sharf, Lee, & Ranaldi, 2005). Previous research has indicated that systemic administration (Broos et al., 2012; Koffarnus et al., 2011; van Gaalen et al., 2006; but see Wade et al., 2000), as well as intramPFC infusions (Loos et al., 2010; but see Yates et al., 2014), of D1 receptor antagonists increases impulsive choice. However, since these previous studies did not determine the effects of D1 receptor blockade on sensitivity to reinforcer magnitude/delayed reinforcement, those results did not address the possibility that D1 receptors are more important for mediating sensitivity to reinforcer amount, as opposed to impulsive choice per se. This is an important distinction because there is some support for the argument that DA signaling in NAcc can be altered by different magnitudes of reinforcement. Specifically, Gan, Walton, and Phillips (2010) observed an increase in dopamine release when they increased the size of the reinforcer. The current results are consistent with this finding, as blocking D1 signaling shifted preference away from the large magnitude reinforcer, even when its delivery was immediate.

An unexpected finding from this study is that the lower concentration of SCH 23390 (0.3 μg) decreased sensitivity to reinforcer magnitude, whereas the higher concentration (1.0 μg) did not. One possible explanation for this finding is that SCH 23390 may have lost its selectivity for D1 antagonism at the higher concentration. For example, in addition to blocking D1 receptors, SCH 23390 inhibits the 5-HT transporter (SERT; Zarrindast, Honardar, Sanea, & Owji, 2011) and acts as an agonist at the 5-HT2C receptor (Ramos, Goñi-Allo, & Aguirre, 2005). Since 5-HT2C receptors are located in the NAcc (Adlersberg et al., 2000) and 5-HT signaling is known to play a role in delay discounting (Miyazaki, Miyazaki, & Doya, 2012; see Homberg, 2012 for a review), the interactive effect of SCH 23390 on both DA and 5-HT systems may explain the apparent biphasic concentration-effect curve observed here.

Regarding the lack of effect of D2 antagonism on discounting in the current study, one important consideration is that D2 receptor antagonists can be mediated by cues that signal the delay to reinforcement. Specifically, Zeeb et al. (2010) reported that a D2 receptor antagonist increased impulsive choice, but only when the delay to delivery of the large magnitude reinforcer was signaled by a stimulus light. One could argue that the null effects observed with eticlopride in the current study may have occurred because we did not signal the delay to delivery of the large reinforcer. However, this seems unlikely, as other research has shown that intracranial infusions of a D2 receptor antagonist into mPFC increases impulsive choice, even though the delay to reinforcement was not directly signaled (Yates et al., 2014). Overall, the current results, in conjunction with past research, suggest that dopaminergic activity in distinct brain regions differentially mediate discounting performance, as D2 receptors in frontal cortical areas may be more important for controlling sensitivity to delayed reinforcement, whereas D1 receptors in NAcc may be more important for controlling sensitivity to reinforcer magnitude.

Whereas blocking D1 receptors decreased sensitivity to reinforcer magnitude, ifenprodil (1.0 μg), which blocks NR2B-containing NMDA receptors, decreased sensitivity to delayed reinforcement. Caution needs to be taken because differences between the highest dose (1.0 μg) of ifenprodil and vehicle only approached statistical significance when all rats were included in the analyses. However, one rat had a b parameter estimate that was 143% higher than the average value following vehicle treatment; additionally, this rat responded for the large reinforcer 60% of the time, even when its delivery was immediate. When this rat was excluded from data analyses, results showed that ifenprodil (1.0 μg) significantly decreased sensitivity to delayed reinforcement. This result is consistent with a previous report showing that Ro 63–1908, a highly selective antagonist for NR2B-containing NMDA receptors, decreases impulsive choice (Higgins et al., 2016). Although the NR2B subunit is not widely expressed in NAcc (Wenzel et al., 1995), administration of neither the uncompetitive NMDA receptor antagonist MK-801 or the competitive antagonist AP-5 altered discounting performance; thus, the current results provide some evidence that the NR2B subunit of the NMDA receptor in NAcc plays a critical role in impulsive choice.

The finding that ifenprodil decreased sensitivity to delayed reinforcement is somewhat at odds with previous work from our laboratory showing that systemic administration of ifenprodil decreases sensitivity to reinforcer magnitude without altering impulsive choice (Yates, Gunkel et al., 2017). However, because ifenprodil binds to other receptors, particularly α1 adrenergic receptors (Chenard et al., 1991), the decreased sensitivity to reinforcer magnitude observed by Yates, Gunkel et al. (2017) could be due to ifenprodil’s actions on adrenergic receptors outside of NAcc. Future work is needed to determine if adrenergic receptors within NAcc differentially mediate sensitivity to reinforcer magnitude/delayed reinforcement, as well as to test the effects of NAcc infusions of highly selective NR2B antagonists, such as Ro 63–1908 or CP-101,606, on discounting.

Although systemic administration of MK-801 has been shown to decrease impulsive choice (Higgins et al., 2016; Yates et al., 2015; but see Yates, Gunkel et al., 2017), NAcc infusions did not significantly alter discounting in our study. This suggests that MK 801-induced alterations in delay discounting are mediated by NMDA receptors outside of NAcc. In addition, since MK-801 increases DA levels in prefrontal cortex (Tsukada et al., 2005), and DA receptor agonists decrease impulsive choice (e.g., Koffarnus et al., 2011), the results obtained by Higgins et al. (2016) and Yates et al. (2015) could be due to the increase in DA in this region as opposed to blocking Glu transmission.

Similar to previous studies (Yates, Gunkel et al., 2017; Yates, Rogers et al., 2017), we showed that the type of analysis used (ANOVA vs. NLME) alters interpretation of the current results. Overall, ANOVA results failed to detect significant differences in discounting following intracranial infusions. However, fitting the exponential discounting function to each subject’s data via NLME allowed us to detect a decrease in sensitivity to reinforcer magnitude following SCH 23390 administration and a decrease in sensitivity to delayed reinforcement (i.e., decreased impulsive choice) following ifenprodil administration. By using the exponential discounting function, we were able to dissociate how receptors within NAcc mediate task performance. In most studies assessing how pharmacological manipulations alter discounting performance, the raw proportion of responses for the large magnitude reinforcer are typically analyzed with ANOVAs (Baarendse & Vanderschuren, 2012; Cardinal, Robbins, & Everitt, 2000; Floresco et al., 2008; Koffarnus et al., 2011; Sukhotina et al., 2008; van Gaalen et al., 2006; Winstanley et al., 2005). Significant effects or interactions are interpreted to reflect either increases or decreases in impulsive choice. However, these interpretations are not always accurate. For example, Yates, Gunkel et al. (2017) recently analyzed the effects of several glutamatergic drugs on delay discounting using ANOVAs and NLME. When ANOVAs were used, the results indicated that ketamine, memantine, and ifenprodil increased impulsive choice; however, NLME analyses indicated that sensitivity to reinforcer magnitude decreased, but not impulsive choice. Future studies should consider how intracranial infusions alter different parameters that may influence an animal’s performance in decision-making procedures.

One procedural limitation to the current study is related to the number of repeated intracranial infusions into the NAcc. Each rat received either nine (DA experiment) or 10 (Glu experiment) infusions. To control for any possible order effects, we counterbalanced the order in which infusions were administered. Because of the small sample sizes (DA experiment: n = 6; Glu experiment: n = 8), we could not conduct statistical analyses to determine if drug infusion order altered the effects of each ligand on discounting, although visual inspection of Tables 1 and 2 did not reveal systematic changes in each parameter estimate as a function of infusion order. Instead, we determined if rats became more sensitive to delayed reinforcement and/or reinforcer magnitude by comparing b and A parameter estimates across each baseline (i.e., the two sessions before each infusion, as well as the two sessions following the final infusion). This analysis showed that neither sensitivity to delayed reinforcement or sensitivity to reinforcer magnitude changed across the experiment. Thus, the alterations in behavior observed following SCH 23390 or ifenprodil administration are not likely to be the result of damage to NAcc.

Another limitation is the use of a discounting procedure in which delays were only increased across the session. Previous research has shown that the order in which delays are presented can modulate drug effects in discounting procedures (Maguire, Henson, & France, 2014; Tanno, Maguire, Henson, & France, 2014; Yates, Rogers et al., 2017). Related to the current results, one could argue that the effects of ifenprodil on discounting may reflect perseverative responding on the lever associated with the large magnitude reinforcer instead of a decrease in impulsive choice. Because we did not include a separate group of rats that were tested on a discounting procedure in which the delays decreased across the session, we cannot rule out this possibility. In a probability discounting procedure, Yates et al. (2016) found that systemic administration of ifenprodil decreased risk-taking behavior when the probabilities of obtaining the large reinforcer increased across the session, but had no effect on behavior when the probabilities decreased across blocks of trials.

Despite these limitations, the results of this study show an apparent dissociation in the role of NAcc DA and Glu receptors on delay discounting. Whereas DA D1 receptors appear to mediate sensitivity to reinforcer magnitude, blocking NR2B-containing NMDA receptors decreases sensitivity to delayed reinforcement (i.e., impulsive choice). Overall, these results provide additional evidence for the utility in applying quantitative analyses to discounting data.

References

Adlersberg, M., Arango, V., Hsiung, S., Mann, J. J., Underwood, M. D., Liu, K., . . .Tamir, H. (2000). In vitro autoradiography of serotonin 5-HT2A/2C receptor-activated G protein: Guanosine-5′-(γ-[35S]thio)triphosphate binding in rat brain. Journal of Neuroscience Research, 61, 674–685. 10.1002/1097-4547(20000915)61:6<674::AID-JNR11>3.0.CO;2-F

Adriani, W., Boyer, F., Gioiosa, L., Macrì, S., Dreyer, J. L., & Laviola, G. (2009). Increased impulsive behavior and risk proneness following lentivirus-mediated dopamine transporter over-expression in rats’ nucleus accumbens. Neuroscience, 159, 47–58.

Baarendse, P. J. J., & Vanderschuren, L. J. M. J. (2012). Dissociable effects of monoamine reuptake inhibitors on distinct forms of impulsive behavior in rats. Psychopharmacology, 219, 313–326. 10.1007/s00213-011-2576-x

Bakshi, V. P., & Geyer, M. A. (1998). Multiple limbic regions mediate the disruption of prepulse inhibition produced in rats by the noncompetitive NMDA antagonist dizocilpine. The Journal of Neuroscience, 18, 8394–8401.

Baldwin, A. E., Holahan, M. R., Sadeghian, K., & Kelley, A. E. (2000). N-methyl-D-aspartate receptor-dependent plasticity within a distributed corticostriatal network mediates appetitive instrumental learning. Behavioral Neuroscience, 114, 84–98. 10.1037/0735-7044.114.1.84

Beninger, R. J., Cheng, M., Hahn, B. L., Hoffman, D. C., Mazurski, E. J., Morency, M. A., . . .Stewart, R. J. (1987). Effects of extinction, pimozide, SCH 23390, and metoclopramide on food-rewarded operant responding of rats. Psychopharmacology, 92, 343–349. 10.1007/BF00210842

Ben-Shahar, O. M., Szumlinski, K. K., Lominac, K. D., Cohen, A., Gordon, E., Ploense, K. L., . . .Woodward, N. (2012). Extended access to cocaine self-administration results in reduced glutamate function within the medial prefrontal cortex. Addiction Biology, 17, 746–757. 10.1111/j.1369-1600.2011.00428.x

Bezzina, G., Body, S., Cheung, T. H. C., Hampson, C. L., Bradshaw, C. M., Szabadi, E., . . .Deakin, J. F. W. (2008). Effect of disconnecting the orbital prefrontal cortex from the nucleus accumbens core on inter-temporal choice behaviour: A quantitative analysis. Behavioural Brain Research, 191, 272–279. 10.1016/j.bbr.2008.03.041

Bezzina, G., Cheung, T. H., Asgari, K., Hampson, C. L., Body, S., Bradshaw, C. M., . . .Anderson, I. M. (2007). Effects of quinolinic acid-induced lesions of the nucleus accumbens core on inter-temporal choice: A quantitative analysis. Psychopharmacology, 195, 71–84. 10.1007/s00213-007-0882-0

Bezzina, G., Cheung, T. H., Body, S., Deakin, J. F., Anderson, I. M., Bradshaw, C. M., & Szabadi, E. (2009). Quantitative analysis of the effect of lesions of the subthalamic nucleus on intertemporal choice: Further evidence for enhancement of the incentive value of food reinforcers. Behavioural Pharmacology, 20, 437–446. 10.1097/FBP.0b013e3283305e4d

Bidwell, L. C., McClernon, F. J., & Kollins, S. H. (2011). Cognitive enhancers for the treatment of ADHD. Pharmacology, Biochemistry, and Behavior, 99, 262–274. 10.1016/j.pbb.2011.05.002

Broos, N., Diergaarde, L., Schoffelmeer, A. N. M., Pattij, T., & De Vries, T. J. (2012). Trait impulsive choice predicts resistance to extinction and propensity to relapse to cocaine seeking: A bidirectional investigation. Neuropsychopharmacology, 37, 1377–1386. 10.1038/npp.2011.323

Cardinal, R. N., Pennicott, D. R., Sugathapala, C. L., Robbins, T. W., & Everitt, B. J. (2001). Impulsive choice induced in rats by lesions of the nucleus accumbens core. Science, 292, 2499–2501. 10.1126/science.1060818

Cardinal, R. N., Robbins, T. W., & Everitt, B. J. (2000). The effects of d-amphetamine, chlordiazepoxide, α-flupenthixol and behavioural manipulations on choice of signalled and unsignalled delayed reinforcement in rats. Psychopharmacology, 152, 362–375. 10.1007/s002130000536

Chenard, B. L., Shalaby, I. A., Koe, B. K., Ronau, R. T., Butler, T. W., Prochniak, M. A., . . .Fox, C. B. (1991). Separation of α 1 adrenergic and N-methyl-D-aspartate antagonist activity in a series of ifenprodil compounds. Journal of Medicinal Chemistry, 34, 3085–3090. 10.1021/jm00114a018

Cohen, J. (1988). Statistical power analysis for the behavioral sciences (2nd ed.). Hillsdale, NJ: Erlbaum.

Cottone, P., Iemolo, A., Narayan, A. R., Kwak, J., Momaney, D., & Sabino, V. (2013). The uncompetitive NMDA receptor antagonists ketamine and memantine preferentially increase the choice for a small, immediate reward in low-impulsive rats. Psychopharmacology, 226, 127–138. 10.1007/s00213-012-2898-3

da Costa Araújo, S., Body, S., Hampson, C. L., Langley, R. W., Deakin, J. F., Anderson, I. M., . . .Szabadi, E. (2009). Effects of lesions of the nucleus accumbens core on inter-temporal choice: Further observations with an adjusting-delay procedure. Behavioural Brain Research, 202, 272–277. 10.1016/j.bbr.2009.04.003

Dar, M. S. (2002). Mouse cerebellar adenosine-glutamate interactions and modulation of ethanol-induced motor incoordination. Alcoholism: Clinical and Experimental Research, 26, 1395–1403.

Denk, F., Walton, M. E., Jennings, K. A., Sharp, T., Rushworth, M. F. S., & Bannerman, D. M. (2005). Differential involvement of serotonin and dopamine systems in cost-benefit decisions about delay or effort. Psychopharmacology, 179, 587–596. 10.1007/s00213-004-2059-4

Dubois, A., Savasta, M., Curet, O., & Scatton, B. (1986). Autoradiographic distribution of the D1 agonist [3H]SKF 38393, in the rat brain and spinal cord. Comparison with the distribution of D2 dopamine receptors. Neuroscience, 19, 125–137. 10.1016/0306-4522(86)90010-2

Evenden, J. L., & Ryan, C. N. (1996). The pharmacology of impulsive behaviour in rats: The effects of drugs on response choice with varying delays of reinforcement. Psychopharmacology, 128, 161–170. 10.1007/s002130050121

Feja, M., Hayn, L., & Koch, M. (2014). Nucleus accumbens core and shell inactivation differentially affects impulsive behaviours in rats. Progress in Neuro-Psychopharmacology & Biological Psychiatry, 54, 31–42. 10.1016/j.pnpbp.2014.04.012

Floresco, S. B., Tse, M. T. L., & Ghods-Sharifi, S. (2008). Dopaminergic and glutamatergic regulation of effort- and delay-based decision making. Neuropsychopharmacology, 33, 1966–1979. 10.1038/sj.npp.1301565

Galtress, T., & Kirkpatrick, K. (2010). The role of the nucleus accumbens core in impulsive choice, timing, and reward processing. Behavioral Neuroscience, 124, 26–43. 10.1037/a0018464

Gan, J. O., Walton, M. E., & Phillips, P. E. M. (2010). Dissociable cost and benefit encoding of future rewards by mesolimbic dopamine. Nature Neuroscience, 13, 25–27. 10.1038/nn.2460

Griffin, W. C., III, Haun, H. L., Hazelbaker, C. L., Ramachandra, V. S., & Becker, H. C. (2014). Increased extracellular glutamate in the nucleus accumbens promotes excessive ethanol drinking in ethanol dependent mice. Neuropsychopharmacology, 39, 707–717. 10.1038/npp.2013.256

Higgins, G. A., Silenieks, L. B., MacMillan, C., Sevo, J., Zeeb, F. D., & Thevarkunnel, S. (2016). Enhanced attention and impulsive action following NMDA receptor GluN2B-selective antagonist pretreatment. Behavioural Brain Research, 311, 1–14. 10.1016/j.bbr.2016.05.025

Hitchcott, P. K., & Phillips, G. D. (1997). Amygdala and hippocampus control dissociable aspects of drug-associated conditioned rewards. Psychopharmacology, 131, 187–195. 10.1007/s002130050283

Ho, M. Y., Mobini, S., Chiang, T. J., Bradshaw, C. M., & Szabadi, E. (1999). Theory and method in the quantitative analysis of “impulsive choice” behaviour: Implications for psychopharmacology. Psychopharmacology, 146, 362–372. 10.1007/PL00005482

Homberg, J. R. (2012). Serotonin and decision making processes. Neuroscience and Biobehavioral Reviews, 36, 218–236. 10.1016/j.neubiorev.2011.06.001

Jensen, V., Rinholm, J. E., Johansen, T. J., Medin, T., Storm-Mathisen, J., Sagvolden, T., . . .Bergersen, L. H. (2009). N-methyl-D-aspartate receptor subunit dysfunction at hippocampal glutamatergic synapses in an animal model of attention-deficit/hyperactivity disorder. Neuroscience, 158, 353–364. 10.1016/j.neuroscience.2008.05.016

Kheramin, S., Body, S., Mobini, S., Ho, M.-Y., Velázquez-Martinez, D. N., Bradshaw, C. M., . . .Anderson, I. M. (2002). Effects of quinolinic acid-induced lesions of the orbital prefrontal cortex on inter-temporal choice: A quantitative analysis. Psychopharmacology, 165, 9–17. 10.1007/s00213-002-1228-6

Koffarnus, M. N., Newman, A. H., Grundt, P., Rice, K. C., & Woods, J. H. (2011). Effects of selective dopaminergic compounds on a delay-discounting task. Behavioural Pharmacology, 22, 300–311. 10.1097/FBP.0b013e3283473bcb

Laurent, V., & Westbrook, R. F. (2008). Distinct contributions of the basolateral amygdala and the medial prefrontal cortex to learning and relearning extinction of context conditioned fear. Learning & Memory, 15, 657–666. 10.1101/lm.1080108

Loos, M., Pattij, T., Janssen, M. C., Counotte, D. S., Schoffelmeer, A. N., Smit, A. B., . . .van Gaalen, M. M. (2010). Dopamine receptor D1/D5 gene expression in the medial prefrontal cortex predicts impulsive choice in rats. Cerebral Cortex, 20, 1064–1070. 10.1093/cercor/bhp167

Maguire, D. R., Henson, C., & France, C. P. (2014). Effects of amphetamine on delay discounting in rats depend upon the manner in which delay is varied. Neuropharmacology, 87, 173–179. 10.1016/j.neuropharm.2014.04.012

Mesches, M. H., Bianchin, M., & McGaugh, J. L. (1996). The effects of intra-amygdala infusion of the AMPA receptor antagonist CNQX on retention performance following aversive training. Neurobiology of Learning and Memory, 66, 324–340. 10.1006/nlme.1996.0073

Miller, E. M., Pomerleau, F., Huettl, P., Gerhardt, G. A., & Glaser, P. E. (2014). Aberrant glutamate signaling in the prefrontal cortex and striatum of the spontaneously hypertensive rat model of attention-deficit/hyperactivity disorder. Psychopharmacology, 231, 3019–3029. 10.1007/s00213-014-3479-4

Miyazaki, K. W., Miyazaki, K., & Doya, K. (2012). Activation of dorsal raphe serotonin neurons is necessary for waiting for delayed rewards. The Journal of Neuroscience, 32, 10451–10457. 10.1523/JNEUROSCI.0915-12.2012

Moschak, T. M., & Mitchell, S. H. (2014). Partial inactivation of nucleus accumbens core decreases delay discounting in rats without affecting sensitivity to delay or magnitude. Behavioural Brain Research, 268, 159–168. 10.1016/j.bbr.2014.03.044

National Research Council. (2011). Guide for the care and use of laboratory animals. Washington, DC: National Academies Press.

Pardey, M. C., Kumar, N. N., Goodchild, A. K., & Cornish, J. L. (2013). Catecholamine receptors differentially mediate impulsive choice in the medial prefrontal and orbitofrontal cortex. Journal of Psychopharmacology, 27, 203–212. 10.1177/0269881112465497

Parkes, S. L., & Balleine, B. W. (2013). Incentive memory: Evidence the basolateral amygdala encodes and the insular cortex retrieves outcome values to guide choice between goal-directed actions. The Journal of Neuroscience, 33, 8753–8763. 10.1523/JNEUROSCI.5071-12.2013

Paxinos, G., & Watson, C. (1998). The rat brain in stereotaxic coordinates (4th ed.). San Diego, CA: Academic Press.

Perlov, E., Philipsen, A., Hesslinger, B., Buechert, M., Ahrendts, J., Feige, B., . . .Tebartz van Elst, L. (2007). Reduced cingulate glutamate/glutamine-to-creatine ratios in adult patients with attention deficit/hyperactivity disorder—A magnet resonance spectroscopy study. Journal of Psychiatric Research, 41, 934–941. 10.1016/j.jpsychires.2006.12.007

Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D., & R Core Team. (2017). nlme: Linear and nonlinear mixed effects models (R package, Version 3.1–131) [Computer software]. Retrieved from https://CRAN.R-project.org/package=nlme

Pothuizen, H. H., Jongen-Rêlo, A. L., Feldon, J., & Yee, B. K. (2005). Double dissociation of the effects of selective nucleus accumbens core and shell lesions on impulsive-choice behaviour and salience learning in rats. European Journal of Neuroscience, 22, 2605–2616. 10.1111/j.1460-9568.2005.04388.x

Ramos, M., Goñi-Allo, B., & Aguirre, N. (2005). Administration of SCH 23390 into the medial prefrontal cortex blocks the expression of MDMA-induced behavioral sensitization in rats: An effect mediated by 5-HT2C receptor stimulation and not by D1 receptor blockade. Neuropsychopharmacology, 30, 2180–2191. 10.1038/sj.npp.1300735

Savasta, M., Dubois, A., & Scatton, B. (1986). Autoradiographic localization of D1 dopamine receptors in the rat brain with [3H]SCH 23390. Brain Research, 375, 291–301. 10.1016/0006-8993(86)90749-3

Sharf, R., Lee, D. Y., & Ranaldi, R. (2005). Microinjections of SCH 23390 in the ventral tegmental area reduce operant responding under a progressive ratio schedule of food reinforcement in rats. Brain Research, 1033, 179–185. 10.1016/j.brainres.2004.11.041

Sombers, L. A., Beyene, M., Carelli, R. M., & Wightman, R. M. (2009). Synaptic overflow of dopamine in the nucleus accumbens arises from neuronal activity in the ventral tegmental area. The Journal of Neuroscience, 29, 1735–1742. 10.1523/JNEUROSCI.5562-08.2009

Sukhotina, I. A., Dravolina, O. A., Novitskaya, Y., Zvartau, E. E., Danysz, W., & Bespalov, A. Y. (2008). Effects of mGlu1 receptor blockade on working memory, time estimation, and impulsivity in rats. Psychopharmacology, 196, 211–220. 10.1007/s00213-007-0953-2

Tanno, T., Maguire, D. R., Henson, C., & France, C. P. (2014). Effects of amphetamine and methylphenidate on delay discounting in rats: Interactions with order of delay presentation. Psychopharmacology, 231, 85–95. 10.1007/s00213-013-3209-3

Tsukada, H., Nishiyama, S., Fukumoto, D., Sato, K., Kakiuchi, T., & Domino, E. F. (2005). Chronic NMDA antagonism impairs working memory, decreases extracellular dopamine, and increases D1 receptor binding in prefrontal cortex of conscious monkeys. Neuropsychopharmacology, 30, 1861–1869. 10.1038/sj.npp.1300732

Valencia-Torres, L., Olarte-Sánchez, C. M., da Costa Araújo, S., Body, S., Bradshaw, C. M., & Szabadi, E. (2012). Nucleus accumbens and delay discounting in rats: Evidence from a new quantitative protocol for analyzing inter-temporal choice. Psychopharmacology, 219, 271–283. 10.1007/s00213-011-2459-1

van Gaalen, M. M., van Koten, R., Schoffelmeer, A. N. M., & Vanderschuren, L. J. M. J. (2006). Critical involvement of dopaminergic neurotransmission in impulsive decision making. Biological Psychiatry, 60, 66–73. 10.1016/j.biopsych.2005.06.005

Wade, T. R., de Wit, H., & Richards, J. B. (2000). Effects of dopaminergic drugs on delayed reward as a measure of impulsive behavior in rats. Psychopharmacology, 150, 90–101. 10.1007/s002130000402

Wenzel, A., Scheurer, L., Künzi, R., Fritschy, J. M., Mohler, H., & Benke, D. (1995). Distribution of NMDA receptor subunit proteins NR2A, 2B, 2C and 2D in rat brain. Neuroreport, 7, 45–48. 10.1097/00001756-199512000-00010

Winstanley, C. A., Theobald, D. E. H., Dalley, J. W., & Robbins, T. W. (2005). Interactions between serotonin and dopamine in the control of impulsive choice in rats: Therapeutic implications for impulse control disorders. Neuropsychopharmacology, 30, 669–682.

Yates, J. R., Batten, S. R., Bardo, M. T., & Beckmann, J. S. (2015). Role of ionotropic glutamate receptors in delay and probability discounting in the rat. Psychopharmacology, 232, 1187–1196. 10.1007/s00213-014-3747-3

Yates, J. R., Breitenstein, K. A., Gunkel, B. T., Hughes, M. N., Johnson, A. B., Rogers, K. K., & Shape, S. M. (2016). Effects of NMDA receptor antagonists on probability discounting depend on the order of probability presentation. Pharmacology, Biochemistry, and Behavior, 150–151, 31–38. 10.1016/j.pbb.2016.09.004

Yates, J. R., Gunkel, B. T., Rogers, K. K., Hughes, M. N., & Prior, N. A. (2017). Effects of N-methyl-D-aspartate receptor ligands on sensitivity to reinforcer magnitude and delayed reinforcement in a delay-discounting procedure. Psychopharmacology, 234, 461–473. 10.1007/s00213-016-4469-5

Yates, J. R., Perry, J. L., Meyer, A. C., Gipson, C. D., Charnigo, R., & Bardo, M. T. (2014). Role of medial prefrontal and orbitofrontal monoamine transporters and receptors in performance in an adjusting delay discounting procedure. Brain Research, 1574, 26–36. 10.1016/j.brainres.2014.06.004

Yates, J. R., Rogers, K. K., Gunkel, B. T., Prior, N. A., Hughes, M. N., Sharpe, S. M., . . .Shults, H. N. (2017). Effects of Group I metabotropic glutamate receptor antagonists on sensitivity to reinforcer magnitude and delayed reinforcement in a delay-discounting task in rats: Contribution of delay presentation order. Behavioural Brain Research, 322, 29–33. 10.1016/j.bbr.2017.01.015

Zarrindast, M. R., Honardar, Z., Sanea, F., & Owji, A. A. (2011). SKF 38393 and SCH 23390 inhibit reuptake of serotonin by rat hypothalamic synaptosomes. Pharmacology, 87, 85–89. 10.1159/000323232

Zeeb, F. D., Floresco, S. B., & Winstanley, C. A. (2010). Contributions of the orbitofrontal cortex to impulsive choice: Interactions with basal levels of impulsivity, dopamine signaling, and reward-related cues. Psychopharmacology, 211, 87–98. 10.1007/s00213-010-1871-2

Zhang, W.-N., Bast, T., & Feldon, J. (2000). Microinfusion of the non-competitive N-methyl-D-aspartate receptor antagonist MK-801 (dizocilpine) into the dorsal hippocampus of Wistar rats does not affect latent inhibition and prepulse inhibition, but increases startle reaction and locomotor activity. Neuroscience, 101, 589–599. 10.1016/S0306-4522(00)00418-8

Submitted: May 17, 2017 Revised: July 6, 2017 Accepted: July 17, 2017

Titel:
Effects of intra-accumbal administration of dopamine and ionotropic glutamate receptor drugs on delay discounting performance in rats.
Autor/in / Beteiligte Person: Yates, JR ; Bardo, MT
Link:
Zeitschrift: Behavioral neuroscience, Jg. 131 (2017-10-01), Heft 5, S. 392
Veröffentlichung: Washington, D.C. : American Psychological Association, [c1983-, 2017
Medientyp: academicJournal
ISSN: 1939-0084 (electronic)
DOI: 10.1037/bne0000214
Schlagwort:
  • 2,3,4,5-Tetrahydro-7,8-dihydroxy-1-phenyl-1H-3-benzazepine
  • Animals
  • Benzazepines
  • Choice Behavior drug effects
  • Dopamine administration & dosage
  • Glutamic Acid metabolism
  • Impulsive Behavior drug effects
  • Male
  • Nucleus Accumbens drug effects
  • Nucleus Accumbens physiology
  • Quinpirole
  • Rats
  • Rats, Sprague-Dawley
  • Receptors, Dopamine D1 antagonists & inhibitors
  • Receptors, Dopamine D1 metabolism
  • Receptors, Dopamine D2 metabolism
  • Receptors, Ionotropic Glutamate drug effects
  • Receptors, N-Methyl-D-Aspartate antagonists & inhibitors
  • Receptors, N-Methyl-D-Aspartate metabolism
  • Reinforcement, Psychology
  • Salicylamides
  • Delay Discounting drug effects
  • Dopamine pharmacology
  • Glutamic Acid pharmacology
Sonstiges:
  • Nachgewiesen in: MEDLINE
  • Sprachen: English
  • Publication Type: Journal Article
  • Language: English
  • [Behav Neurosci] 2017 Oct; Vol. 131 (5), pp. 392-405.
  • MeSH Terms: Delay Discounting / *drug effects ; Dopamine / *pharmacology ; Glutamic Acid / *pharmacology ; 2,3,4,5-Tetrahydro-7,8-dihydroxy-1-phenyl-1H-3-benzazepine ; Animals ; Benzazepines ; Choice Behavior / drug effects ; Dopamine / administration & dosage ; Glutamic Acid / metabolism ; Impulsive Behavior / drug effects ; Male ; Nucleus Accumbens / drug effects ; Nucleus Accumbens / physiology ; Quinpirole ; Rats ; Rats, Sprague-Dawley ; Receptors, Dopamine D1 / antagonists & inhibitors ; Receptors, Dopamine D1 / metabolism ; Receptors, Dopamine D2 / metabolism ; Receptors, Ionotropic Glutamate / drug effects ; Receptors, N-Methyl-D-Aspartate / antagonists & inhibitors ; Receptors, N-Methyl-D-Aspartate / metabolism ; Reinforcement, Psychology ; Salicylamides
  • References: Neuroscience. 1986 Sep;19(1):125-37. (PMID: 2946980) ; Neuroscience. 2009 Jan 12;158(1):353-64. (PMID: 18571865) ; J Neurosci Res. 2000 Sep 15;61(6):674-85. (PMID: 10972964) ; Behav Brain Res. 2008 Aug 22;191(2):272-9. (PMID: 18472170) ; Psychopharmacology (Berl). 2017 Feb;234(3):461-473. (PMID: 27837332) ; Psychopharmacology (Berl). 2012 Jan;219(2):313-26. (PMID: 22134476) ; Psychopharmacology (Berl). 1999 Oct;146(4):362-72. (PMID: 10550487) ; Psychopharmacology (Berl). 1996 Nov;128(2):161-70. (PMID: 8956377) ; Psychopharmacology (Berl). 2000 Nov;152(4):362-75. (PMID: 11140328) ; Neuroscience. 2000;101(3):589-99. (PMID: 11113308) ; J Neurosci. 2012 Aug 1;32(31):10451-7. (PMID: 22855794) ; Learn Mem. 2008 Aug 26;15(9):657-66. (PMID: 18772253) ; Science. 2001 Jun 29;292(5526):2499-501. (PMID: 11375482) ; Brain Res. 1986 Jun 11;375(2):291-301. (PMID: 2942221) ; Psychopharmacology (Berl). 2007 Nov;195(1):71-84. (PMID: 17659381) ; J Psychiatr Res. 2007 Dec;41(11):934-41. (PMID: 17303167) ; Biol Psychiatry. 2006 Jul 1;60(1):66-73. (PMID: 16125144) ; Behav Brain Res. 2017 Mar 30;322(Pt A):29-33. (PMID: 28088471) ; Behav Neurosci. 2000 Feb;114(1):84-98. (PMID: 10718264) ; Neurosci Biobehav Rev. 2012 Jan;36(1):218-36. (PMID: 21693132) ; Psychopharmacology (Berl). 2013 Mar;226(1):127-38. (PMID: 23104264) ; Psychopharmacology (Berl). 2012 Jan;219(2):271-83. (PMID: 21894486) ; J Psychopharmacol. 2013 Feb;27(2):203-12. (PMID: 23135240) ; Nat Neurosci. 2010 Jan;13(1):25-7. (PMID: 19904261) ; Neuropsychopharmacology. 2012 May;37(6):1377-86. (PMID: 22318198) ; Neuropsychopharmacology. 2005 Apr;30(4):669-82. (PMID: 15688093) ; Neuropsychopharmacology. 2005 Dec;30(12):2180-91. (PMID: 15841107) ; Behav Neurosci. 2010 Feb;124(1):26-43. (PMID: 20141278) ; J Neurosci. 2009 Feb 11;29(6):1735-42. (PMID: 19211880) ; Neuropharmacology. 2014 Dec;87:173-9. (PMID: 24780379) ; Psychopharmacology (Berl). 2014 Aug;231(15):3019-29. (PMID: 24682500) ; Neurobiol Learn Mem. 1996 Nov;66(3):324-40. (PMID: 8946425) ; Psychopharmacology (Berl). 2002 Dec;165(1):9-17. (PMID: 12474113) ; J Neurosci Methods. 1980 Dec;3(2):129-49. (PMID: 6110810) ; Neuroreport. 1995 Dec 29;7(1):45-8. (PMID: 8742413) ; Neuropsychopharmacology. 2005 Oct;30(10):1861-9. (PMID: 15841110) ; Addict Biol. 2012 Jul;17(4):746-57. (PMID: 22339852) ; J Med Chem. 1991 Oct;34(10):3085-90. (PMID: 1681106) ; Behav Brain Res. 2009 Sep 14;202(2):272-7. (PMID: 19463712) ; J Neurosci. 1998 Oct 15;18(20):8394-401. (PMID: 9763482) ; Psychopharmacology (Berl). 2005 May;179(3):587-96. (PMID: 15864561) ; Neuropsychopharmacology. 2014 Feb;39(3):707-17. (PMID: 24067300) ; Psychopharmacology (Berl). 2015 Apr;232(7):1187-96. (PMID: 25270726) ; Behav Pharmacol. 2011 Aug;22(4):300-11. (PMID: 21694584) ; Psychopharmacology (Berl). 2008 Feb;196(2):211-20. (PMID: 17909752) ; Psychopharmacology (Berl). 2010 Jul;211(1):87-98. (PMID: 20428999) ; Brain Res. 2005 Feb 8;1033(2):179-85. (PMID: 15694922) ; J Neurosci. 2013 May 15;33(20):8753-63. (PMID: 23678118) ; Psychopharmacology (Berl). 1987;92(3):343-9. (PMID: 2957718) ; Brain Res. 2014 Jul 29;1574:26-36. (PMID: 24928616) ; Eur J Neurosci. 2005 Nov;22(10):2605-16. (PMID: 16307603) ; Psychopharmacology (Berl). 2014 Jan;231(1):85-95. (PMID: 23963529) ; Behav Brain Res. 2016 Sep 15;311:1-14. (PMID: 27180168) ; Behav Pharmacol. 2009 Sep;20(5-6):437-46. (PMID: 19667971) ; Psychopharmacology (Berl). 2000 May;150(1):90-101. (PMID: 10867981) ; Behav Brain Res. 2014 Jul 15;268:159-68. (PMID: 24704637) ; Neuropsychopharmacology. 2008 Jul;33(8):1966-79. (PMID: 17805307) ; Pharmacol Biochem Behav. 2016 Nov - Dec;150-151:31-38. (PMID: 27642050) ; Pharmacology. 2011;87(1-2):85-9. (PMID: 21242715) ; Prog Neuropsychopharmacol Biol Psychiatry. 2014 Oct 3;54:31-42. (PMID: 24810333) ; Cereb Cortex. 2010 May;20(5):1064-70. (PMID: 19690230) ; Psychopharmacology (Berl). 1997 May;131(2):187-95. (PMID: 9201808) ; Pharmacol Biochem Behav. 2011 Aug;99(2):262-74. (PMID: 21596055)
  • Grant Information: P50 DA005312 United States DA NIDA NIH HHS; T32 DA007304 United States DA NIDA NIH HHS; T32 DA016176 United States DA NIDA NIH HHS; United States NH NIH HHS
  • Substance Nomenclature: 0 (Benzazepines) ; 0 (Receptors, Dopamine D1) ; 0 (Receptors, Dopamine D2) ; 0 (Receptors, Ionotropic Glutamate) ; 0 (Receptors, N-Methyl-D-Aspartate) ; 0 (SCH 23390) ; 0 (Salicylamides) ; 20OP60125T (Quinpirole) ; 3KX376GY7L (Glutamic Acid) ; 67287-49-4 (2,3,4,5-Tetrahydro-7,8-dihydroxy-1-phenyl-1H-3-benzazepine) ; J8M468HBH4 (eticlopride) ; VTD58H1Z2X (Dopamine)
  • Entry Date(s): Date Created: 20170929 Date Completed: 20180522 Latest Revision: 20240610
  • Update Code: 20240610
  • PubMed Central ID: PMC5679283

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -