Neurobiologesch Substrate fir d'Däischter Säit vu Compulsivitéit an der Sucht (2009)

Neuropharmacology. Author manuscript; available in PMC Jan 1, 2010.

PMCID: PMC2637927

NIHMSID: NIHMS86836

Den Editeur seng lescht geännert Versioun vun dësem Artikel ass verfügbar Neuropharmakologie

Kuckt aner Artikelen an PMC dat ze verëffentlechen den publizéierten Artikel.

Géi op:

mythologesch

Drug addiction can be defined by a compulsion to seek and take drug, loss of control in limiting intake, and the emergence of a negative emotional state when access to the drug is prevented. Drug addiction impacts multiple motivational mechanisms and can be conceptualized as a disorder that progresses from impulsivity (positive reinforcement) to compulsivity (negative reinforcement). The construct of negative reinforcement is defined as drug taking that alleviates a negative emotional state. The negative emotional state that drives such negative reinforcement is hypothesized to derive from dysregulation of key neurochemical elements involved in reward and stress within the basal forebrain structures involving the ventral striatum and extended amygdala. Specific neurochemical elements in these structures include not only decreases in reward neurotransmission, such as decreases in dopamine and opioid peptide function in the ventral striatum, but also recruitment of brain stress systems, such as corticotropin-releasing factor (CRF), in the extended amygdala. Acute withdrawal from all major drugs of abuse produces increases in reward thresholds, increases in anxiety-like responses, and increases in extracellular levels of CRF in the central nucleus of the amygdala. CRF receptor antagonists also block excessive drug intake produced by dependence. A brain stress response system is hypothesized to be activated by acute excessive drug intake, to be sensitized during repeated withdrawal, to persist into protracted abstinence, and to contribute to the compulsivity of addiction. Other components of brain stress systems in the extended amygdala that interact with CRF and may contribute to the negative motivational state of withdrawal include norepinephrine, dynorphin, and neuropeptide Y. The combination of loss of reward function and recruitment of brain stress systems provides a powerful neurochemical basis for a negative emotional state that is responsible for the negative reinforcement driving, at least in part, the compulsivity of addiction.

Schlësselwieder: Sucht, Oppositiounsprozess, Stress, erweidert Amygdala, Kortikotropin-Releasedfaktor

1. Definitions and conceptual framework for compulsivity in addiction

Drug addiction is a chronically relapsing disorder characterized by (i) compulsion to seek and take the drug, (ii) loss of control in limiting intake, and (iii) emergence of a negative emotional state (e.g., dysphoria, anxiety, irritability) reflecting a motivational withdrawal syndrome when access to the drug is prevented (defined here as dependence) (Koob a Le Moal, 1997). Sucht gëtt ugeholl datt si mat hirem Syndrom identesch ass Substanz Ofhängegkeet (wéi aktuell definéiert vun der Diagnostesch a statistesch Handbuch vu Mentalenstudenten; American Psychiatric Association, 1994). Clinically and in animal models, the occasional but limited use of a drug with the Potential fir Mëssbrauch oder Ofhängegkeet ass ënnerschiddlech vun Epilatioun an Drogenënnerscheed ze ginn an d'Entstehung vun enger chronescher Drogenabhängiger Staat.

Drug addiction has been conceptualized as a disorder that involves elements of both impulsivity and compulsivity, where Impulsivitéit can be defined behaviorally as “a predisposition toward rapid, unplanned reactions to internal and external stimuli without regard for the negative consequences of these reactions to themselves or others” (Moeller et al., 2001). Impulsivity is measured in two domains: the choice of a smaller, immediate reward over a larger, delayed reward (Rachlin and Green, 1972) or the inability to inhibit behavior by changing the course of action or to stop a response once it is initiated (Logan et al., 1997). Impulsivity is a core deficit in substance abuse disorders (Allen et al., 1998) and in neuropsychiatric disorders such as attention deficit hyperactivity disorder. Operationally, delay-to-gratification tasks (delayed discounting tasks) (impulsive choice) and the stop-signal or go/no-go task (behavioral impulsivity) have been used as measures of impulsivity (Fillmore and Rush, 2002; Green et al., 1994). Compulsivitéit can be defined as elements of behavior that result in perseveration in responding in the face of adverse consequences or perseveration in the face of incorrect responses in choice situations. These elements are analogous to the symptoms of Substance Dependence as outlined by the American Psychiatric Association: continued substance use despite knowledge of having had a persistent or recurrent physical or psychological problem and a great deal of time spent in activities necessary to obtain the substance (Amerikanescher Psychiatrie Associatioun, 2000).

Collapsing the cycles of impulsivity and compulsivity yields a composite addiction cycle comprised of three stages—bezéien / anticipatioun, binge / wäerch, an Auszug / negativ Auswierkung—where impulsivity often dominates at the early stages and compulsivity dominates at terminal stages. As an individual moves from impulsivity to compulsivity, a shift occurs from positive reinforcement driving the motivated behavior to negative reinforcement driving the motivated behavior (Koob, 2004). Negativ Verstäerkung kann definéiert ginn als de Prozess, duerch deen d'Entfernung vun engem aversive Stimulus (zB negativen emotionalen Zoustand vum Medikamententzug) d'Wahrscheinlechkeet vun enger Äntwert erhéicht (zB Ofhängegkeet-induzéiert Medikamenter). Dës dräi Etappe sinn konzeptualiséiert wéi mateneen interagéieren, méi intensiv ginn, a schlussendlech zum pathologesche Staat féieren bekannt als Sucht (Koob a Le Moal, 1997) (Table 1). The present review will focus on the role of an animal model of compulsivity that derives from the negative emotional state of the Auszug / negativ Auswierkung stage of the addiction cycle.

Table 1 

Etappen vum Suchtzyklus.

Different drugs produce different patterns of addiction with emphasis on different components of the addiction cycle. Opioids can be considered classic drugs of addiction because subjects meet most of the criteria classically associated with addiction, including dramatic tolerance and withdrawal. A pattern of intravenous or smoked drug-taking evolves, including intense intoxication, the development of tolerance, escalation in intake, and profound dysphoria, physical discomfort, and somatic withdrawal signs during abstinence. Intense preoccupation with obtaining opioids (craving) develops that often precedes the somatic signs of withdrawal and is linked not only to stimuli associated with obtaining the drug but also to stimuli associated with withdrawal and the aversive motivational state. A pattern develops where the drug must be obtained to avoid the severe dysphoria and discomfort of abstinence. Other drugs of abuse follow a similar pattern but may involve more the binge / wäerch stage (psychostimulants and alcohol) or less binge / wäerch a méi Auszug / negativ Auswierkung an bezéien / anticipatioun stages (nicotine and cannabinoids).

Alcohol addiction, or alcoholism, can follow a similar trajectory, but the pattern of oral drug taking is often characterized by binges of alcohol intake that can be daily episodes or prolonged days of heavy drinking and is characterized by a severe emotional and somatic withdrawal syndrome. Many alcoholics continue with such a binge/withdrawal pattern for extended periods, but some individuals can evolve into an opioid-like situation in which they must have alcohol available at all times to avoid the negative consequences of abstinence. Tobacco addiction contrasts with the above patterns—the binge / wäerch stage forms a minor component of nicotine dependence. The pattern of nicotine intake is one of highly titrated intake of the drug except during periods of sleep. However, during abstinence, users experience negative emotional states, including dysphoria, irritability, and intense craving. Marijuana Dependence follows a pattern similar to opioids and tobacco, with a significant intoxication stage, but as chronic use continues, subjects begin to show a pattern of use manifest by chronic intoxication during waking hours and withdrawal characterized by dysphoria, irritability, and sleep disturbances. Psychostimulants, such as cocaine and amphetamines, show a pattern focused on the binge / wäerch stage in which binges can be hours or days in duration and often are followed by a withdrawal (“crash”) characterized by extreme dysphoria and inactivity.

1.1. Motivation, withdrawal, and opponent process

Motivation is a state that can be defined as a “tendency of the whole animal to produce organized activity” (Hebb, 1972), and such motivational states are not constant but rather vary over time. Early work by Wikler stressed the role of changes in drive states associated with dependence. Subjects described withdrawal changes as a “hunger” or primary need and the effects of morphine on such a state as “satiation” or gratification of the primary need (Wikler, 1952). Although Wikler argued that positive reinforcement was retained even in heavily dependent subjects (thrill of the intravenous opioid injection), dependence produced a new source of gratification, that of negative reinforcement (see above).

The concept of motivation was linked inextricably with hedonic, affective, or emotional states in addiction in the context of temporal dynamics by Solomon’s opponent process theory of motivation. Solomon and Corbit (1974) postulated that hedonic, affective, or emotional states, once initiated, are automatically modulated by the central nervous system with mechanisms that reduce the intensity of hedonic feelings. The a-process includes affective or hedonic habituation (or tolerance), and the b-Prozess includes affective or hedonic withdrawal (abstinence). The a-process in drug use consists of positive hedonic responses, occurs shortly after presentation of a stimulus, correlates closely with the intensity, quality, and duration of the reinforcer, and shows tolerance. In contrast, the b-Prozess in drug use appears after the a-process has terminated, consists of negative hedonic responses, and is sluggish in onset, slow to build up to an asymptote, slow to decay, and gets larger with repeated exposure. The thesis here is that opponent processes begin early in drug-taking, reflect changes in the brain reward and stress systems, and later form one of the major motivations for compulsivity in drug taking.

Thus, dependence or manifestation of a withdrawal syndrome after removal of chronic drug administration is defined in terms of motivéiert aspects of dependence such as emergence of a negative emotional state (e.g., dysphoria, anxiety, irritability) when access to the drug is prevented (Koob a Le Moal, 2001), rather than on the kierperlech signs of dependence. Indeed, some have argued that the development of such a negative affective state can define dependence as it relates to addiction:

“The notion of dependence on a drug, object, role, activity or any other stimulus-source requires the crucial feature of negative affect experienced in its absence. The degree of dependence can be equated with the amount of this negative affect, which may range from mild discomfort to extreme distress, or it may be equated with the amount of difficulty or effort required to do without the drug, object, etc” (Russell, 1976).

Rapid acute tolerance and opponent process-like effects in response to the hedonic effects of cocaine have been reported in human studies of smoked coca paste (Van Dyke and Byck, 1982) (Figure 1A). After a single smoking session, the onset and intensity of the “high” are very rapid via the smoked route of administration, and a rapid tolerance is manifest. The “high” decreases rapidly despite significant blood levels of cocaine. Even more intriguing is that human subjects also actually report a subsequent “dysphoria,” again despite high blood levels of cocaine. Intravenous cocaine produced similar patterns of a rapid “rush” followed by an increased “low” in human laboratory studies (Breiter et al., 1997) (Figure 1B). With intravenous cocaine self-administration in animal models, such elevations in reward threshold begin rapidly and can be observed within a single session of self-administration (Kenny et al., 2003) (Figure 2), bearing a striking resemblance to human subjective reports. These results demonstrate that the elevation in brain reward thresholds following prolonged access to cocaine failed to return to baseline levels between repeated, prolonged exposure to cocaine self-administration (i.e., residual hysteresis), thus creating a greater and greater elevation in “baseline” ICSS thresholds. These data provide compelling evidence for brain reward dysfunction in escalated cocaine self-administration that provide strong support for a hedonic allostasis model of drug addiction.

Figure 1Figure 1 

(A) Dysphoric feelings followed the initial euphoria in experimental subjects who smoked cocaine paste, even though the concentration of cocaine in the plasma of the blood remained relatively high. The dysphoria is characterized by anxiety, depression, ...
Figure 2 

Rat (n = 11) were allowed to self-administer 10, 20, 40, and 80 injections of cocaine (0.25 mg per injection), and intracranial self-stimulation reward thresholds were measured 15 min and 2, 24, and 48 h after the end of each intravenous cocaine self-administration ...

Similar results have been observed showing dysphoria-like responses accompanying acute opioid and ethanol withdrawal (Liu and Schulteis, 2004; Schulteis and Liu, 2006). Here, naloxone administration following single injections of morphine increased reward thresholds, measured by ICSS, and increased thresholds with repeated morphine and naloxone-induced withdrawal experience (Liu and Schulteis, 2004). Similar results were observed during repeated acute withdrawal from ethanol (Schulteis and Liu, 2006).

The dysregulation of brain reward function associated with withdrawal from chronic administration of drugs of abuse is a common element of all drugs of abuse. Withdrawal from chronic cocaine (Markou a Koob, 1991), Amphetamin (Paterson et al., 2000), opioids (Schulteis et al., 1994), cannabinoids (Gardner a Virel, 1998), Nikotin (Epping-Jordan et al., 1998), and ethanol (Schulteis et al., 1995) leads to increases in reward threshold during acute abstinence, and some of these elevations in threshold can last for up to one week (Figure 3). These observations lend credence to the hypothesis that opponent processes can set the stage for one aspect of compulsivity where negative reinforcement mechanisms are engaged.

Figure 3 

(A) Effects of ethanol withdrawal on CRF-like immunoreactivity (CRF-L-IR) in the rat amygdala determined by microdialysis. Dialysate was collected over four 2 h periods regularly alternated with nonsampling 2 h periods. The four sampling periods corresponded ...

More recently, opponent process theory has been expanded into the domains of the neurobiology of drug addiction from a neurocircuitry perspective. An allostatic model of the brain motivational systems has been proposed to explain the persistent changes in motivation that are associated with dependence in addiction (Koob a Le Moal 2001, 2008). In this formulation, addiction is conceptualized as a cycle of increasing dysregulation of brain reward/anti-reward mechanisms that results in a negative emotional state contributing to the compulsive use of drugs. Counteradaptive processes that are part of the normal homeostatic limitation of reward function fail to return within the normal homeostatic range. These counteradaptive processes are hypothesized to be mediated by two mechanisms: within-system neuroadaptations and between-system neuroadaptations (Koob a Bloom, 1988).

In a within-system neuroadaptation, “the primary cellular response element to the drug would itself adapt to neutralize the drug’s effects; persistence of the opposing effects after the drug disappears would produce the withdrawal response” (Koob a Bloom, 1988). Thus, a within-system neuroadaptation is a molecular or cellular change within a given reward circuit to accommodate overactivity of hedonic processing associated with addiction resulting in a decrease in reward function.

The emotional dysregulation associated with the Auszug / negativ Auswierkung stage also may involve between-system neuroadaptations in which neurochemical systems other than those involved in the positive rewarding effects of drugs of abuse are recruited or dysregulated by chronic activation of the reward system (Koob a Bloom, 1988). Thus, a between-system neuroadaptation is a circuitry change in which another different circuit (anti-reward circuit) is activated by the reward circuit and has opposing actions, again limiting reward function. The purpose of this review is to explore the neuroadaptational changes that occur in the brain emotional systems to account for the neurocircuitry changes that produce opponent processes and are hypothesized to have a key role in the compulsivity of addiction.

1.2. Animal models of compulsivity in addiction measured by negative emotional-like states: Place aversion, animal models of anxiety, and reward thresholds

Animal models of the Auszug / negativ Auswierkung stage include measures of conditioned place aversion (rather than preference) to precipitated withdrawal or spontaneous withdrawal from chronic administration of a drug, increases in reward thresholds using brain stimulation reward (Markou a Koob, 1991; Schulteis et al., 1994, 1995; Epping-Jordan et al., 1998; Gardner a Virel, 1998; Paterson et al., 2000), and increases in anxiety-like responses (for review, see Shippenberg a Koob, 2002; Sanchis-Segura and Spanagel, 2006).

1.3. Animal models of compulsivity in addiction as defined by increased drug taking: Escalation in drug self-administration with prolonged access

A progressive increase in the frequency and intensity of drug use is one of the major behavioral phenomena characterizing the development of addiction and has face validity with the criteria of the Diagnostesch a statistesch Handbuch vu Mentalenstudenten: “The substance is often taken in larger amounts and over a longer period than was intended” (American Psychological Association, 1994). A framework with which to model the transition from drug use to drug addiction can be found in recent animal models of prolonged access to intravenous cocaine self-administration. Historically, animal models of cocaine self-administration involved the establishment of stable behavior from day to day to allow the reliable interpretation of data provided by within-subject designs aimed at exploring the neuropharmacological and neurobiological bases of the reinforcing effects of acute cocaine. Up until 1998, after acquisition of self-administration, rats typically were allowed access to cocaine for 3 h or less per day to establish highly stable levels of intake and patterns of responding between daily sessions. This was a useful paradigm for exploring the neurobiological substrates for the acute reinforcing effects of drugs of abuse.

However, in an effort to explore the possibility that differential access to intravenous cocaine self-administration in rats may produce different patterns of drug intake, rats were allowed access to intravenously self-administration cocaine for 1 or 6 h per day (Ahmed a Koob, 1998). One hour access (short access or ShA) to intravenous cocaine per session produced low and stable intake as observed previously. In contrast, 6 h access (long access or LgA) to cocaine produced drug intake that gradually escalated over days (Figure 4). Increased intake was observed in the extended access group during the first hour of the session, with sustained intake over the entire session and an upward shift in the dose-effect function, suggesting an increase in hedonic set point. When animals were allowed access to different doses of cocaine, both the LgA and ShA animals titrated their cocaine intake, but the LgA rats consistently self-administered almost twice as much cocaine at any dose tested, further suggesting an upward shift in the set point for cocaine reward in the escalated animals (Ahmed a Koob, 1999; Deroche-Gamonet et al., 2004; Mantsch et al., 2004). Escalation also is associated with an increase in break point for cocaine in a progressive-ratio schedule of reinforcement, suggesting an enhanced motivation to seek cocaine or an enhanced efficacy of cocaine reward (Paterson an Markou, 2003; Wee et al., 2008). Such increased self-administration in dependent animals has now been observed with cocaine, methamphetamine, nicotine, heroin, and alcohol (Ahmed et al., 2000; Ahmed a Koob, 1998; Kitamura et al., 2006; O'Dell et al., 2004; George et al., 2007) (Figure 4). This model is a key element for evaluating the motivational significance of opponent process changes in the brain reward and stress systems in addiction that lead to compulsivity in addiction. Similar changes in the reinforcing and incentive effects of cocaine as drug intake have been observed following extended access and include increased cocaine-induced reinstatement after extinction and a decreased latency to goal time in a runway model for cocaine reward (Deroche et al., 1999). Altogether, these results suggest that drug taking with extended access changes the motivation to seek the drug. Whether this enhanced drug taking reflects a sensitization of reward or a reward deficit state remains under discussion (Vezina, 2004), but the brain reward and neuropharmacological studies outlined below argue for a reward deficit state driving the increased drug taking during extended access.

Figure 4 

(A) Effect of drug availability on cocaine intake (mean ± SEM). In 6 h long-access (LgA) rats (n = 12) but not in 1 h short-access (ShA) rats (n = 12), mean total cocaine intake started to increase significantly from session 5 (p <0.05; ...

The hypothesis that compulsive cocaine use is accompanied by a chronic perturbation in brain reward homeostasis has been tested in an animal model of escalation in drug intake with prolonged access combined with measures of brain stimulation reward thresholds. Animals implanted with intravenous catheters and allowed differential access to intravenous self-administration of cocaine showed increases in cocaine self-administration from day to day in the long-access group (6 h; LgA) but not in the short-access group (1 h; ShA). The differential exposure to cocaine self-administration had dramatic effects on reward thresholds that progressively increased in LgA rats but not in ShA or control rats across successive self-administration sessions (Ahmed et al., 2002). Elevation in baseline reward thresholds temporally preceded and was highly correlated with escalation in cocaine intake (Figure 5). Post-session elevations in reward thresholds failed to return to baseline levels before the onset of each subsequent self-administration session, thereby deviating more and more from control levels. The progressive elevation in reward thresholds was associated with the dramatic escalation in cocaine consumption that was observed previously. After escalation had occurred, an acute cocaine challenge facilitated brain reward responsiveness to the same degree as before but resulted in higher absolute brain reward thresholds in LgA compared with ShA rats (Ahmed et al., 2002). Similar results have been observed with extended access to heroin (Kenny et al., 2006). Rats allowed 23 h access to heroin also showed a time-dependent increase in reward thresholds that paralleled the increases in heroin intake (Figure 5).

Figure 5Figure 5 

(A) Relationship between elevation in intracranial self-stimulation reward thresholds and cocaine intake escalation. (Left) Percentage change from baseline ICSS thresholds. (Right) Number of cocaine injections earned during the first hour of each session. ...

2. Neural substrates for the negative emotional state associated with addiction

2.1. Within-system neuroadaptations that contribute to the negative emotional state component of compulsivity

Elektresch Gehirnstimulatioun Belounung oder intracranial Self-Stimulatioun huet eng laang Geschicht wéi e Mesure vun Aktivitéit vum Gehirer Belounungssystem a vun de akuten Verstärkungseffekter vun Drogen vu Mëssbrauch. All Medikamenter vu Mëssbrauch, wéi séier vergewëssert, reduzéieren d'Gehirnstimulatioun Belounungsschwellen (Kornetsky and Esposito, 1979) and when administered chronically increase reward thresholds during withdrawal (see above). Brain stimulation reward involves widespread neurocircuitry in the brain, but the most sensitive sites defined by the lowest thresholds involve the trajectory of the medial forebrain bundle that connects the ventral tegmental area with the basal forebrain (Olds an Milner, 1954; Koob et al., 1977). While much emphasis was focused initially on the role of the ascending monoamine systems in the medial forebrain bundle, other nondopaminergic systems in the medial forebrain bundle clearly have a key role (Hernandez et al., 2006).

Within-system neuroadaptations to chronic drug exposure include decreases in function of the same neurotransmitter systems in the same neurocircuits implicated in the acute reinforcing effects of drugs of abuse. One prominent hypothesis is that dopamine systems are compromised in crucial phases of the addiction cycle, such as withdrawal and leads to decreased motivation for non-drug-related stimuli and increased sensitivity to the abused drug (Melis et al., 2005). Activation of the mesolimbic dopamine system has long been known to be critical for the acute rewarding properties of psychostimulant drugs and to be associated with the acute reinforcing effects of other drugs of abuse (Koob, 1992; Di Chiara and North, 1992; Nestler, 2005). Decreases in activity of the mesolimbic dopamine system and decreases in serotonergic neurotransmission in the nucleus accumbens occur during drug withdrawal in animal studies (Rossetti et al., 1992; Weiss et al., 1992, 1996). Imaging studies in drug-addicted humans have consistently shown long-lasting decreases in the numbers of dopamine D2 Rezeptoren bei Drogenhéierern verglach mat Kontrollen (Volkow et al., 2002). In addition, cocaine abusers have reduced dopamine release in response to a pharmacological challenge with a stimulant drug (Volkow et al., 1997; Martinez et al., 2007). Decreases in the number of dopamine D2 receptors, coupled with the decrease in dopaminergic activity, in cocaine, nicotine, and alcohol abusers results in decreased sensitivity of reward circuits to stimulation by natural reinforcers (Martin-Solch et al., 2001; Volkow a Fowler, 2000). These findings suggest an overall reduction in the sensitivity of the dopamine component of reward circuitry to natural reinforcers and other drugs in drug-addicted individuals.

Psychostimulant withdrawal in humans is associated with fatigue, decreased mood, and psychomotor retardation and in animals is associated with decreased motivation to work for natural rewards (Barr an Phillips, 1999) a reduzéierter Lokomotoraktivitéit (Pulvirenti and Koob, 1993), Verhalenseffekter déi verréngert dopaminergesch Funktioun involvéiere kënnen. Déieren während Amphetamin Réckzuch weisen ofgeholl Äntwert op engem progressive Verhältnis Zäitplang fir eng séiss Léisung, an dës ofgeholl Äntwert gouf ëmgedréint vun der Dopamin partiell agonist Terguride (Orsini et al., 2001), suggesting that low dopamine tone contributes to the motivational deficits associated with psychostimulant withdrawal.

Under this conceptual framework, other within-system neuroadaptations would include increased sensitivity of receptor transduction mechanisms in the nucleus accumbens. Drugs of abuse have acute receptor actions that are linked to intracellular signaling pathways that may undergo adaptations with chronic treatment. Activation of adenylate cyclase, protein kinase A, cyclic adenosine monophosphate response-element binding protein (CREB), and ΔFosB has been observed during opioid withdrawal (Self et al., 1995; Shaw-Lutchman et al., 2002; Nye an Nestler, 1996; Nestler, 2004). The ΔFosB response is hypothesized to represent a neuroadaptive change that extends long into protracted abstinence (Nestler and Malenka, 2004).

2.2. Between-system neuroadaptations that contribute to the negative emotional state component of compulsivity

Brain neurochemical systems involved in arousal-stress modulation also may be engaged within the neurocircuitry of the brain stress systems in an attempt to overcome the chronic presence of the perturbing drug and to restore normal function despite the presence of drug. Both the hypothalamic-pituitary-adrenal axis and the brain stress system mediated by corticotropin-releasing factor (CRF) are dysregulated by chronic administration of all major drugs with dependence or abuse potential, with a common response of elevated adrenocorticotropic hormone, corticosterone, and amygdala CRF during acute withdrawal (Rivier et al., 1984; Merlo-Pich et al., 1995; Koob et al., 1994; Rasmussen et al., 2000; Olive et al., 2002; Delfs et al., 2000). Acute withdrawal from all drugs of abuse also produces an aversive or anxiety-like state that can be reversed by CRF antagonists (see below).

Den neuroanatomeschen Entity bezeechent den erweiderten Amygdal (Heimer an Alheid, 1991) may represent a common anatomical substrate integrating brain arousal-stress systems with hedonic processing systems to produce the between-system opponent process elaborated above. The extended amygdala is composed of the central nucleus of the amygdala, bed nucleus of the stria terminalis, and a transition zone in the medial (shell) subregion of the nucleus accumbens. Each of these regions has cytoarchitectural and circuitry similarities (Heimer an Alheid, 1991). D'erweidmete Amyldaala kritt vill Afferen aus limbesche Strukturen wéi de basolateral Ammdala an Hippocampus a schéckt Äfter op de mediale Deel vum ventralen Pallidum an eng grouss Préjatioun mat dem lateralen Hypothalamus, sou datt d'spezifesch Gehirberezonen definéiert sinn, déi d'klassesch Limbik (emotional) Strukturen mat dem extrapyramidalen Motorsystem (Alheid et al., 1995). Déi verlängert Amygdal huet laang laang hypothesch gemaach, fir eng Schlësselroll ze hunn net nëmmen an Angscht konditionnéieren (Le Doux, 2000), awer och am emotionalen Deel vun der Schmerzveraarbechtung (Neugebauer et al., 2004).

2.3. Neuropharmacological studies of the aversive stimulus effects of drug withdrawal

Place aversion has been used to measure the aversive stimulus effects of withdrawal, mostly in the context of opioids (Hand et al., 1988; Stinus et al., 1990). In contrast to conditioned place preference, rats exposed to a particular environment while undergoing precipitated withdrawal to opioids spend less time in the withdrawal-paired environment when subsequently presented with a choice between that environment and an unpaired environment. These aversive stimulus effects can be measured from 24 h to 16 weeks later (Hand et al., 1988; Stinus et al., 1990, 2000). The place aversion does not require maintenance of opioid dependence for its manifestation. Such an association continues to be manifested weeks after animals are “detoxified” (e.g., after the morphine pellets are removed) (see Baldwin and Koob, 1993; Stinus et al., 2000). In addition, a place aversion in opioid-dependent rats can be observed with doses of naloxone below which somatic signs of withdrawal are observed (Schulteis et al., 1994). Although naloxone itself will produce a place aversion in nondependent rats, the threshold dose required to produce a place aversion decreases significantly in dependent rats (Hand et al., 1988). A variant on this approach is to explore the place aversion produced following naloxone injection after a single acute injection of morphine.

Acute opioid dependence has been defined as the precipitation of withdrawal-like signs by opioid antagonists following a single dose or short-term administration of an opioid agonist (Martin and Eades, 1964). Rats show a reliable conditioned place aversion precipitated by a low dose of naloxone after a single morphine injection that reflects a motivational component of acute withdrawal (Azar et al., 2003). Acute opioid withdrawal also produces increases in reward thresholds (Liu and Schulteis, 2004), suppression in operant responding (Schulteis et al., 2003) and increased anxiety-like behavior in the elevated plus maze (Zhang and Schulteis, 2008). Using the conditioned place aversion paradigm, the opioid partial agonist buprenorphine dose-dependently decreased the place aversion produced by precipitated opioid withdrawal. Systemic administration of a CRF1 Rezeptor Antagonist an direkt intracerebral Verwaltung vun engem Peptid CRF1/ CRF2 Antagonist huet och opioid Réckzuch-induzéiert Plaz Aversionen reduzéiert (Stinus et al., 2005; Heinrichs et al., 1995). Functional noradrenergic antagonists blocked opioid withdrawal-induced place aversion (Delfs et al., 2000).

Another candidate for the aversive effects of drug withdrawal is dynorphin. Much evidence shows that dynorphin is increased in the nucleus accumbens in response to dopaminergic activation and, in turn, that overactivity of the dynorphin systems can decrease dopaminergic function. κ opioid agonists are aversive (Land et al., 2008; Pfeiffer et al., 1986), and cocaine, opioid, and ethanol withdrawal is associated with increased dynorphin in the nucleus accumbens and/or amygdala (Spangler et al., 1993; Lindholm et al., 2000; Rattan et al., 1992).

2.4. Neuropharmacological studies of the anxiety-like effects of drug withdrawal

Another common response to acute withdrawal and protracted abstinence from all major drugs of abuse is the manifestation of anxiety-like responses. Animal models have revealed anxiety-like response to all major drugs of abuse during acute withdrawal. The dependent variable is often a passive response to a novel and/or aversive stimulus, such as the open field or elevated plus maze, or an active response to an aversive stimulus, such as defensive burying of an electrified metal probe. Withdrawal from repeated administration of cocaine produces an anxiogenic-like response in the elevated plus maze and defensive burying test, both of which are reversed by administration of CRF antagonists (Sarnyai et al., 1995; Basso et al., 1999). Ausgefall Réckzuch an opioid Ofhängegkeet produzéiert och Angscht-ähnlechen Effekter (Schulteis et al., 1998; Harris an Aston Jones, 1993). Ethanol withdrawal produces anxiety-like behavior that is reversed by intracerebroventricular administration of CRF1/ CRF2 peptidergesche Antagonisten (Baldwin et al., 1991), small molecule CRF1 Antagonist (Knapp et al., 2004; Overstreet et al., 2004; Funk et al., 2007), and intracerebral administration of a peptidergic CRF1/ CRF2 antagonist an d'Amygdala (Rassnick et al., 1993). CRF antagonists injected intracerebroventricularly or systemically also block the potentiated anxiety-like responses to stressors observed during protracted abstinence from chronic ethanol (Breese et al., 2005; Valdez et al., 2003). The effects of CRF antagonists have been localized to the central nucleus of the amygdala (Rassnick et al., 1993). Ausgefällt Réckzuch vum Nikotin produzéiert ängschtlech Äntwerten déi och vun CRF Antagonisten ëmgedréit ginn (Tucci et al., 2003; George et al., 2007).

3. Neural Substrates for Increased Drug-taking with Extended Access

3.1. Within-system neuroadaptations

In a series of studies, dopamine partial agonists have not only been shown to reverse psychostimulant withdrawal but also to block the increase in psychostimulant self-administration associated with extended access. Dopamine partial agonists decrease the reinforcing effects of psychostimulant drugs in nondependent limited access paradigms (Izzo et al., 2001; Pulvirenti et al., 1998). However, animals with extended access show an increased sensitivity to a dopamine partial agonist (Wee et al., 2007). Long-access rats administered a dopamine D2 partial agonist showed a shift to the left of the dose-response function similar to results observed with dopamine antagonists (Ahmed a Koob, 2004). These results, combined with the observation that dopamine partial agonists also can reverse psychostimulant withdrawal, suggest that dysregulation of dopamine tone may contribute to the motivational effects of drug withdrawal.

Similar results have been observed in opioid dependence with the opioid partial agonist buprenorphine. Buprenorphine dose-dependently decreased heroin self-administration in opioid-dependent rats (Chen et al., 2006).

3.2. Between-system neuroadaptations

The ability of CRF antagonists to block the anxiogenic-like and aversive-like motivational effects of drug withdrawal would predict motivational effects of CRF antagonists in animal models of extended access to drugs. CRF antagonists selectively blocked the increased self-administration of drugs associated with extended access to intravenous self-administration of cocaine (Specio et al., 2008), Nikotin (George et al., 2007) an Heroin (Greenwell et al., 2008a). CRF-Antagonisten blockéiert och déi verstäerkte Selbstverwaltung vun Ethanol bei ofhängege Ratten (Funk et al., 2007) (Table 2).

Table 2 

Role of corticotropin-releasing factor in dependence.

Administration of CRF1 antagonists systemically reversed the increased self-administration of cocaine associated with extended access, and this reversal was at antagonist doses that were lower than those that decreased short-access self-administration (Specio et al., 2008). Here, rats allowed 6 h access to cocaine showed an increase in cocaine intake over time, whereas 1 h access rats remained stable. Two different CRF1 antagonists blocked cocaine self-administration in the long-access rats at doses lower than those that blocked cocaine self-administration in short-access rats (Specio et al., 2008).

As noted above, CRF in parts of the extended amygdala is involved in the aversive stimulus effects of opioid withdrawal. The selective CRF1 antagonist antalarmin blocked the place aversion produced by naloxone in morphine-dependent rats (Stinus et al., 2005). CRF1 Noutfallmätscher konnten net bedingte Konditiounsplang ze gesinn fir opsiidt Récktrëtt a keng Opioid-induzéiert Zuel vu Dynorphin mRNA am Nukleus accumbens ze weisen (Contarino a Papaleo, 2005). CRF1 antagonists also selectively blocked the increase in heroin self-administration observed in heroin-dependent rats with extended access (Greenwell et al., 2008a).

Also, as noted above, precipitated withdrawal from chronic nicotine produced anxiogenic-like effects that were blocked by a CRF1 Rezeptor Antagonist (George et al., 2007) and increases in reward thresholds that were reversed by a CRF antagonist (Bruijnzeel et al., 2007). Extracellular CRF has been shown to be increased in the amygdala during withdrawal from chronic nicotine (George et al., 2007). From a developmental perspective, increased CRF-like immunoreactivity has been observed in adult rats exposed to nicotine during adolescence and has been linked to an anxiety-like phenotype (Slawicki et al., 2005). Systemic administration of a CRF1 antagonist blocked the increased self-administration of nicotine associated with withdrawal in extended access (23 h) animals (George et al., 2007). These results suggest that CRF in the basal forebrain also may have an important role in the development of the aversive motivational effects that drive the increased drug seeking associated with cocaine, heroin, and nicotine dependence.

E ganz dramateschen Beispill vun den motivationalen Effekter vum CRF bei der Ofhängegkeet kann an Déiermodeller vun der Ethanol-Selbstverwaltung an onselbstännegen Déieren beobachtet ginn. Während Ethanol Ofleef ginn déi extrahypothalamesch CRF-Systemen hyperaktiv, mat enger Erhéijung vun extrazellulärer CRF innerhalb den zentrale Keel vum Amygdala an de Bettkäre vum Stria-Terminalis vun onselbstabilte Ratten (Funk et al., 2006; Merlo-Pich et al., 1995; Olive et al., 2002). The dysregulation of brain CRF systems is hypothesized to underlie not only the enhanced anxiety-like behaviors but also the enhanced ethanol self-administration associated with ethanol withdrawal. Supporting this hypothesis, the -helical CRF9-41 an D-Phe CRF12-41 (intracerebroventricular administration) reduce ethanol self-administration in dependent animals (Valdez et al., 2004). Exposure to repeated cycles of chronic ethanol vapor produced substantial increases in ethanol intake in rats both during acute withdrawal and during protracted abstinence (2 weeks post-acute withdrawal) (O'Dell et al., 2004; Rimondini et al., 2002). Intracerebroventricular administration of a CRF1/ CRF2 antagonist blocked the dependence-induced increase in ethanol self-administration during both acute withdrawal and protracted abstinence (Valdez et al., 2004). When administered directly into the central nucleus of the amygdala, a CRF1/ CRF2 antagonist blocked ethanol self-administration in ethanol-dependent rats (Funk et al., 2006, 2007). Systemin Injektiounen vu klenge Molekül CRF1 antagonists also blocked the increased ethanol intake associated with acute withdrawal (Knapp et al., 2004; Overstreet et al., 2004; Funk et al., 2007). Dës Donnéeën proposéieren eng wichteg Roll fir CRF, virun allem am Zentralkierper vun der Amygdal, bei der Vermëttlung vun der verstäerkter Selbstverwaltung, déi mat der Ofhängegkeet ass.

Although less well developed, evidence exists for a role of norepinephrine systems in the extended amygdala in the negative motivational state and increased self-administration associated with dependence. Norepinephrine functional antagonists (β1 Antagonist an α2 agonistesche) Injektioun an de Querbettkern vun der Stria Terminalis blockéiert d'Opiate vu Récktrëtt indirekte Plaz Aversiounen (Delfs et al., 2000). The effects of norepinephrine in mediating the motivational effects of opioid withdrawal involve the ventral noradrenergic system. Ventral noradrenergic bundle lesions attenuated opioid withdrawal-induced place aversions (Delfs et al., 2000), but virtually complete lesions of the dorsal noradrenergic bundle from the locus coeruleus with the neurotoxin 6-hydroxydopamine failed to block the place aversion produced by opioid withdrawal (Caille et al., 1999). Functional norepinephrine antagonists block excessive drug intake associated with dependence on ethanol (Walker et al., 2008), Kokain (Wee et al., 2008) an Opioiden (Greenwell et al., 2008b). A focal point for many of these effects is the extended amygdala but at the level of the bed nucleus of the stria terminalis. κ Dynorphin, an opioid peptide that binds to κ opioid receptors, has long been known to show activation with chronic administration of psychostimulants and opioids (Nestler, 2004; Koob, 2008), and κ opioid agonists produce aversive effects in animals and humans (Mucha a Herz, 1985; Pfeiffer et al., 1986). A κ opioid antagonist blocks the excessive drinking associated with ethanol withdrawal and dependence (Walker a Koob, 2008). Evidence demonstrates that κ receptor activation can produce CRF release (Lidd a Takemori, 1992), mä vläicht e puer hunn argumentéiert datt d'Auswirkungen vun Dynorphin bei der Erzeegung vun negative emotional Zustands duerch Aktivatioun vu CRF Systeme vermittelt ginn (Land et al., 2008).

The dynamic nature of the brain stress system response to challenge is illustrated by the pronounced interaction of central nervous system CRF systems and central nervous system norepinephrine systems. Conceptualized as a feed-forward system at multiple levels of the pons and basal forebrain, CRF activates norepinephrine, and norepinephrine in turn activates CRF (Koob, 1999). Vill pharmakologesch, physiologesch an anatomesch Beweiser ënnerstëtzen eng wichteg Roll fir eng CRF-Norepinephrin Interaktioun an der Regioun vum locus coeruleus als Äntwert op Stressoren (Valentino et al., 1991, 1993; Van Bockstaele et al., 1998). Wéi och ëmmer, Noradrenalin stimuléiert och CRF Verëffentlechung am paraventrikuläre Kär vum Hypothalamus (Alonso et al., 1986), the bed nucleus of the stria terminalis, and the central nucleus of the amygdala. Such feed-forward systems were further hypothesized to have powerful functional significance for mobilizing an organism’s response to environmental challenge, but such a mechanism may be particularly vulnerable to pathology (Koob, 1999).

Neuropeptide Y (NPY) is a neuropeptide with dramatic anxiolytic-like properties localized to the amygdala and has been hypothesized to have opposite effects to CRF in the negative motivational state of withdrawal from drugs of abuse (Heilig a Koob, 2007). Significant evidence suggests that activation of NPY in the central nucleus of the amygdala can block the motivational aspects of dependence associated with chronic ethanol administration. NPY administered intracerebroventricularly blocked the increased drug intake associated with ethanol dependence (Thorsell et al., 2005a, b). Injektioun vu NPY direkt an den Zentralkierper vun der Amygdal (Gilpin et al., 2008) and viral vector-enhanced expression of NPY in the central nucleus of the amygdala also blocked the increased drug intake associated with ethanol dependence (Thorsell et al., 2007).

Dofir steigt den akuten Entzug vu Medikamente CRF an den Zentralkierper vun der Amygdala, déi motivativ Bedeitung fir d'Ängstlech-ähnlech Effekter vum akute Récktrëtt an d'erhéijen Drogeproblemer ass mat der Ofhängegkeet ass (Figure 6). Acute withdrawal also may increase the release of norepinephrine in the bed nucleus of the stria terminalis and dynorphin in the nucleus accumbens, and both of which may contribute to the negative emotional state associated with dependence. Decreased activity of NPY in the central nucleus of the amygdala also may contribute to the anxiety-like state associated with ethanol dependence. Activation of brain stress systems (CRF, norepinephrine, dynorphin) combined with inactivation of brain anti-stress systems (NPY) elicits powerful emotional dysregulation in the extended amygdala. Such dysregulation of emotional processing may be a significant contribution to the between-system opponent processes that help maintain dependence and also set the stage for more prolonged state changes in emotionality such as protracted abstinence.

Figure 6 

Neurocircuitry assoziéiert mat den akuten positiven Verstäerkungseffekter vun Drogen vu Mëssbrauch an der negativer Verstäerkung vun der Ofhängegkeet a wéi et ännert sech am Iwwergang vun onofhängege Medikament huelen an ofhängeg Medikament huelen. Schlësselelementer vun der Belounung ...

4. Compulsivity in addiction: An allostatic view

Compulsivity in addiction can derive from multiple sources, including enhanced incentive salience, engagement of habit function, and impairment in executive function. However, underlying each of these sources is a negative emotional state that may strongly impact on compulsivity. The development of the negative emotional state that drives the negative reinforcement of addiction has been defined as the “dark side” of addiction (Koob a Le Moal, 2005, 2008) and is hypothesized to be the b-Prozess of the hedonic dynamic known as opponent process when the a-process is euphoria. The negative emotional state that comprises the Auszug / negativ Auswierkung stage consists of key motivational elements, such as chronic irritability, emotional pain, malaise, dysphoria, alexithymia, and loss of motivation for natural rewards, and is characterized in animals by increases in reward thresholds during withdrawal from all major drugs of abuse. Two processes are hypothesized to form the neurobiological basis for the b-Prozess: loss of function in the reward systems (within-system neuroadaptation) and recruitment of the brain stress or anti-reward systems (between-system neuroadaptation) (Koob a Bloom, 1988; Koob a Le Moal, 1997). Anti-Belounung ass e Konstrukt mat der Hypothese baséiert datt Gehirersysteme sinn an der Plaz fir d'Belaaschtung ze limitéieren (Koob a Le Moal, 2008). As dependence and withdrawal develop, brain stress systems such as CRF, norepinephrine, and dynorphin are recruited, producing aversive or stress-like states (Koob, 2003; Nestler, 2001; Aston-Jones et al., 1999). At the same time, within the motivational circuits of the ventral striatum-extended amygdala, reward function decreases. The combination of decreases in reward neurotransmitter function and recruitment of anti-reward systems provides a powerful source of negative reinforcement that contributes to compulsive drug-seeking behavior and addiction (Figure 6).

An overall conceptual theme argued here is that drug addiction represents a break with homeostatic brain regulatory mechanisms that regulate the emotional state of the animal. The dysregulation of emotion begins with the binge and subsequent acute withdrawal, but leaves a residual neuroadaptive trace that allows rapid “re-addiction” even months and years after detoxification and abstinence. Thus, the emotional dysregulation of drug addiction represents more than simply a homeostatic dysregulation of hedonic function; it also represents a dynamic break with homeostasis of this system that has been termed allostasis.

Allostasis, originally conceptualized to explain persistent morbidity of arousal and autonomic function, is defined as “stability through change.” Allostasis is far more complex than homeostasis and has several unique characteristics (Sterling an Eyer, 1988). Allostasis involves a feed-forward mechanism rather than the negative feedback mechanisms of homeostasis, with continuous re-evaluation of need and continuous readjustment of all parameters toward new set points. An allostatesche Staat kann definéiert ginn als Zoustand vun der chronescher Ofwécklung vum Regulatiounssystem vun sengem normalen (homo-stateschen) Betribsniveau. Allostatesch Laascht was defined as the “long-term cost of allostasis that accumulates over time and reflects the accumulation of damage that can lead to pathological states” (McEwen, 2000).

Allostatic mechanisms have been hypothesized to be involved in maintaining a functioning brain reward system that has relevance for the pathology of addiction (Koob a Le Moal, 2001). Two components are hypothesized to adjust to challenges to the brain produced by drugs of abuse: overactivation of brain reward transmitters and circuits and recruitment of the brain anti-reward or brain stress systems. Thus, the very physiological mechanism that allows rapid responses to environmental challenge becomes the source of pathology if adequate time or resources are not available to shut off the response (e.g., the interaction between CRF and norepinephrine in the brainstem and basal forebrain that could lead to pathological anxiety) (Koob, 1999).

Repeated challenges, such as with drugs of abuse, lead to attempts of the brain via molecular, cellular, and neurocircuitry changes to maintain stability but at a cost. For the drug addiction framework elaborated here, the residual deviation from normal brain reward threshold regulation is termed an allostatesche Staat. This state represents a combination of chronic elevation of reward set point fueled by decreased function of reward circuits and recruitment of anti-reward systems, both of which lead to the compulsivity of drug-seeking and drug-taking. How these systems are modulated by other known brain emotional systems localized to the basal forebrain, where the ventral striatum and extended amygdala project to convey emotional valence, how the dysregulation of brain emotional systems impact on the cognitive domain linked to impairments in executive function, and how individuals differ at the molecular-genetic level of analysis to convey loading on these circuits remain challenges for future research.

Arbeschterlidder

The author would like to thank Michael Arends and Mellany Santos for their outstanding assistance with the preparation of this manuscript. Research was supported by National Institutes of Health grants AA06420 and AA08459 from the National Institute on Alcohol Abuse and Alcoholism, DA10072, DA04043, and DA04398 from the National Institute on Drug Abuse, and DK26741 from the National Institute of Diabetes and Digestive and Kidney Diseases. Research also was supported by the Pearson Center for Alcoholism and Addiction Research. This is publication number 19480 from The Scripps Research Institute.

Noten

Verzichterklärung vum Verlag: Dëst ass eng PDF-Datei vun engem net verworfene Manuskript deen fir d'Publikatioun akzeptéiert gouf. Als Service fir eis Cliente këmmeren mir dës fréie Versioun vum Manuskript. De Manuskript wäert d'Kopieveraarbechtung, d'Setzgarantie an d'Iwwerpréiwung vum entstinnende Beweis ënnerwerfen, ier se an der Final citat Form publizéiert gëtt. Maacht weg datt während dem Produktiounsfehler eventuell Décisiounen entdeckt kënne ginn, déi d'Inhalter beaflosse kënnen, an all gesetzlech Verännerungen, déi op d'Zäitschrëft gelidden sinn.

Referenze

  1. Ahmed SH, Kenny PJ, Koob GF, Markou A. Neurobiological evidence for hedonic allostasis associated with escalating cocaine use. Nature Neuroscience. 2002;5:625–626. [PubMed]
  2. Ahmed SH, Koob GF. Iwwergang vu moderne bis iwwerdrohend Drogeproblemer: Ännerung am hedonesche Sollwert. Wëssenschaft. 1998; 282: 298-300. [PubMed]
  3. Ahmed SH, Koob GF. Long-lasting increase in the set point for cocaine self-administration after escalation in rats. Psychopharmacology. 1999;146:303–312. [PubMed]
  4. Ahmed SH, Koob GF. Changes in response to a dopamine antagonist in rats with escalating cocaine intake. Psychopharmacology. 2004;172:450–454. [PubMed]
  5. Ahmed SH, Walker JR, Koob GF. Persistent Erhéijung vun der Motivatioun fir Heroin an Ratten mat enger Geschicht vun der Drogen-Équilatioun ze huelen. Neuropsychopharmacologie. 2000; 22: 413-421. [PubMed]
  6. Alheid GF, De Olmos JS, Beltramino CA. Amygdala and extended amygdala. In: Paxinos G, editor. The Rat Nervous System. Academic Press; San Diego: 1995. pp. 495–578.
  7. Allen TJ, Moeller FG, Rhoades HM, Cherek DR. Impulsivity and history of drug dependence. Drug and Alcohol Dependence. 1998;50:137–145. [PubMed]
  8. Alonso G, Szafarczyk A, Balmefrezol M, Assenmacher I. Immunocytochemical evidence for stimulatory control by the ventral noradrenergic bundle of parvocellular neurons of the paraventricular nucleus secreting corticotropin-releasing hormone and vasopressin in rats. Brain Research. 1986;397:297–307. [PubMed]
  9. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders. 4. American Psychiatric Press; Washington DC: 2000.
  10. Aston-Jones G, Delfs JM, Druhan J, Zhu Y. The bed nucleus of the stria terminalis: a target site for noradrenergic actions in opiate withdrawal. In: McGinty JF, editor. Advancing from the Ventral Striatum to the Extended Amygdala: Implications for Neuropsychiatry and Drug Abuse. New York Academy of Sciences; New York: 1999. pp. 486–498. (series title: Annals of the New York Academy of Sciences, vol 877) [PubMed]
  11. Azar MR, Jones BC, Schulteis G. Conditioned place aversion is a highly sensitive index of acute opioid dependence and withdrawal. Psychopharmacology. 2003;170:42–50. [PubMed]
  12. Baldwin HA, Koob GF. Rapid induction of conditioned opiate withdrawal in the rat. Neuropsychopharmacology. 1993;8:15–21. [PubMed]
  13. Baldwin HA, Rassnick S, Rivier J, Koob GF, Britton KT. CRF antagonist reverses the “anxiogenic” response to ethanol withdrawal in the rat. Psychopharmacology. 1991;103:227–232. [PubMed]
  14. Barr AM, Phillips AG. Withdrawal following repeated exposure to d-amphetamine decreases responding for a sucrose solution as measured by a progressive ratio schedule of reinforcement. Psychopharmacology. 1999;141:99–106. [PubMed]
  15. Basso AM, Spina M, Rivier J, Vale W, Koob GF. Corticotropin-releasing factor antagonist attenuates the “anxiogenic-like” effect in the defensive burying paradigm but not in the elevated plus-maze following chronic cocaine in rats. Psychopharmacology. 1999;145:21–30. [PubMed]
  16. Breese GR, Overstreet DH, Knapp DJ, Navarro M. Viru méi Ethanol ofgehale ginn d'Stress indukt Angst-ähnlech Verhalens: d'Inhibierung vum CRF1- a Benzodiazepin-Rezeptor-Antagonisten an e 5-HT1aReceptor agonist. Neuropsychopharmacologie. 2005; 30: 1662-1669. [PMC gratis Artikel] [PubMed]
  17. Breiter HC, Gollub RL, Weisskoff RM, Kennedy DN, Makris N, Berke JD, Goodman JM, Kantor HL, Gastfriend DR, Riorden JP, Mathew RT, Rosen BR, Hyman SE. Acute effects of cocaine on human brain activity and emotion. Neuron. 1997;19:591–611. [PubMed]
  18. Bruijnzeel AW, Marcinkiewcz C, Isaac S, Booth MM, Dennis DM, Gold MS. The effects of buprenorphine on fentanyl withdrawal in rats. Psychopharmacology. 2007;191:931–941. [PubMed]
  19. Caille S, Espejo EF, Reneric JP, Cador M, Koob GF, Stinus L. Total neurochemical lesion of noradrenergic neurons of the locus coeruleus does not alter either naloxone-precipitated or spontaneous opiate withdrawal nor does it influence ability of clonidine to reverse opiate withdrawal. Journal of Pharmacology and Experimental Therapeutics. 1999;290:881–892. [PubMed]
  20. Chen SA, O’Dell L, Hoefer M, Greenwell TN, Zorrilla EP, Koob GF. Unlimited access to heroin self-administration: independent motivational markers of opiate dependence. Neuropsychopharmacology. 2006;31:2692–2707. [corrigedum: 31, 2802] [PubMed]
  21. Contarino A, Papaleo F. The corticotropin-releasing factor receptor-1 pathway mediates the negative affective states of opiate withdrawal. Proceedings of the National Academy of Sciences USA. 2005;102:18649–18654. [PMC gratis Artikel] [PubMed]
  22. Delphi JM, Zhu Y, Druhan JP, Aston-Jones G. Noradrenalin am ventralen Forebrain ass kritesch fir opiate Réckgang induzéiert Aversion. Natur. 2000; 403: 430-434. [PubMed]
  23. Deroche V, Le Moal M, Piazza PV. Cocaine self-administration increases the incentive motivational properties of the drug in rats. European Journal of Neuroscience. 1999;11:2731–2736. [PubMed]
  24. Deroche-Gamonet V, Belin D, Piazza PV. Beweiser fir Sucht-ähnlech Verhalen an der Rass. Wëssenschaft. 2004; 305: 1014-1017. [PubMed]
  25. Di Chiara G, North RA. Neurobiology of opiate abuse. Trends in Pharmacological Sciences. 1992;13:185–193. [PubMed]
  26. Epping-Jordanien MP, Watkins SS, Koob GF, Markou A. Dramatesch Verloschter am Gehirer Belounungsfunktion während Nikotinentrackung. Natur. 1998; 393: 76-79. [PubMed]
  27. Fillmore MT, Rush CR. Impaired inhibitory control of behavior in chronic cocaine users. Drug and Alcohol Dependence. 2002;66:265–273. [PubMed]
  28. Funk CK, O’Dell LE, Crawford EF, Koob GF. Corticotropin-releasing factor within the central nucleus of the amygdala mediates enhanced ethanol self-administration in withdrawn, ethanol-dependent rats. Journal of Neuroscience. 2006;26:11324–11332. [PubMed]
  29. Funk CK, Zorrilla EP, Lee MJ, Rice KC, Koob GF. Corticotropin-releasing factor 1 antagonists selectively reduce ethanol self-administration in ethanol-dependent rats. Biological Psychiatry. 2007;61:78–86. [PMC gratis Artikel] [PubMed]
  30. Gardner EL, Vorel SR. Cannabinoid transmission and reward-related events. Neurobiology of Disease. 1998;5:502–533. [PubMed]
  31. George O, Ghozland S, Azar MR, Cottone P, Zorrilla EP, Parsons LH, O'Dell LE, Richardson HN, Koob GF. CRF-CRF1 system activation mediates withdrawal-induced increases in nicotine self-administration in nicotine-dependent rats. Proceedings of the National Academy of Sciences USA. 2007;104:17198–17203. [PMC gratis Artikel] [PubMed]
  32. Gilpin NW, Richardson HN, Lumeng L, Koob GF. Dependence-induced alcohol drinking by alcohol-preferring (P) rats and outbred Wistar rats. Alcoholism: Clinical and Experimental Research; 2008. (in press) [PMC gratis Artikel] [PubMed]
  33. Green L, Fristoe N, Myerson J. Temporal discounting and preference reversals in choice between delayed outcomes. Psychonomic Bulletin and Review. 1994;1:383–389. [PubMed]
  34. Greenwell TN, Funk CK, Cottone P, Richardson HN, Chen SA, Rice K, Lee MJ, Zorrilla EP, Koob GF. Corticotropin-releasing factor-1 receptor antagonists decrease heroin self-administration in long-, but not short-access rats. Addiction Biology. 2008a in press. [PMC gratis Artikel] [PubMed]
  35. Gréngwell TN, Walker BM, Cottone P, Zorrilla EP, Koob GF. Den α1 adrenergic receptor antagonist prazosin reduces heroin self-administration in rats with extended access to heroin administration. Pharmacology Biochemistry and Behavior. 2008b in press. [PMC gratis Artikel] [PubMed]
  36. Hand TH, Koob GF, Stinus L, Le Moal M. Aversive properties of opiate receptor blockade: Evidence for exclusively central mediation in naive and morphine-dependent rats. Brain Research. 1988;474:364–368. [PubMed]
  37. Harris GC, Aston-Jones G. β-Adrenergic antagonists attenuate withdrawal anxiety in cocaine- and morphine-dependent rats. Psychopharmacology. 1993;113:131–136. [PubMed]
  38. Hebb DO. Textbook of Psychology. 3. W.B. Saunders; Philadelphia: 1972.
  39. Heilig M, Koob GF. A key role for corticotropin-releasing factor in alcohol dependence. Trends in Neurosciences. 2007;30:399–406. [PMC gratis Artikel] [PubMed]
  40. Heimer L, Alheid G. Piecing together the puzzle of basal forebrain anatomy. In: Napier TC, Kalivas PW, Hanin I, editors. The Basal Forebrain: Anatomy to Function. Plenum Press; New York: 1991. pp. 1–42. (series title: Advances in Experimental Medicine and Biology, vol 295)
  41. Heinrichs SC, Menzaghi F, Schulteis G, Koob GF, Stinus L. Suppression of corticotropin-releasing factor in the amygdala attenuates aversive consequences of morphine withdrawal. Behavioural Pharmacology. 1995;6:74–80. [PubMed]
  42. Hernandez G, Hamdani S, Rajabi H, Conover K, Stewart J, Arvanitogiannis A, Shizgal P. Prolonged rewarding stimulation of the rat medial forebrain bundle: neurochemical and behavioral consequences. Behavioral Neuroscience. 2006;120:888–904. [PubMed]
  43. Izzo E, Orsini C, Koob GF, Pulvirenti L. A dopamine partial agonist and antagonist block amphetamine self-administration in a progressive ratio schedule. Pharmacology Biochemistry and Behavior. 2001;68:701–708. [PubMed]
  44. Kenny PJ, Chen SA, Kitamura O, Markou A, Koob GF. Conditioned withdrawal drives heroin consumption and decreases reward sensitivity. Journal of Neuroscience. 2006;26:5894–5900. [PubMed]
  45. Kenny PJ, Polis I, Koob GF, Markou A. Low dose cocaine self-administration transiently increases but high dose cocaine persistently decreases brain reward function in rats. European Journal of Neuroscience. 2003;17:191–195. [PubMed]
  46. Kitamura O, Wee S, Specio SE, Koob GF, Pulvirenti L. Escalation of methamphetamine self-administration in rats: a dose-effect function. Psychopharmacology. 2006;186:48–53. [PubMed]
  47. Knapp DJ, Overstreet DH, Moy SS, Breese GR. SB242084, Flumazenil, a CRA1000-Block Ethanol-Récktrëtt-Angscht Angscht an Rattern. Alkohol. 2004; 32: 101-111. [PMC gratis Artikel] [PubMed]
  48. Koob GF. Drugs of abuse: anatomy, pharmacology, and function of reward pathways. Trends in Pharmacological Sciences. 1992;13:177–184. [PubMed]
  49. Koob GF. Corticotropin-releasing factor, norepinephrine and stress. Biological Psychiatry. 1999;46:1167–1180. [PubMed]
  50. Koob GF. Neuroadaptive mechanisms of addiction: studies on the extended amygdala. European Neuropsychopharmacology. 2003;13:442–452. [PubMed]
  51. Koob GF. Allostatic view of motivation: implications for psychopathology. In: Bevins RA, Bardo MT, editors. Motivational Factors in the Etiology of Drug Abuse. University of Nebraska Press; Lincoln NE: 2004. pp. 1–18. (series title: Nebraska Symposium on Motivation, vol 50)
  52. Koob GF. Eng Roll fir Stresssystem vun Hirnniveau an der Sucht. Neuron. 2008; 59: 11-34. [PMC gratis Artikel] [PubMed]
  53. Koob GF, Bloom FE. Cellulare a molekulare Mechanismen vun der Drogenabhängegkeet. Wëssenschaft. 1988; 242: 715-723. [PubMed]
  54. Koob GF, Heinrichs SC, Menzaghi F, Pich EM, Britton KT. Corticotropin releasing factor, stress and behavior. Seminars in the Neurosciences. 1994;6:221–229.
  55. Koob GF, Le Moal M. Drogenmiessung: hedonesch homo-statesch Dysregulatioun. Wëssenschaft. 1997; 278: 52-58. [PubMed]
  56. Koob GF, Le Moal M. Drogen-Sucht, Dysregulatioun vun Belounung an Allostasis. Neuropsychopharmacologie. 2001; 24: 97-129. [PubMed]
  57. Koob GF, Le Moal M. Plasticity of reward neurocircuitry and the “dark side” of drug addiction. Nature Neuroscience. 2005;8:1442–1444. [PubMed]
  58. Koob GF, Le Moal M. Addiction and the brain antireward system. Annual Review of Psychology. 2008;59:29–53. [PubMed]
  59. Koob GF, Winger GD, Meyerhoff JL, Annau Z. Effects of d-amphetamine on concurrent self-stimulation of forebrain and brain stem loci. Brain Research. 1977;137:109–126. [PubMed]
  60. Kornetsky C, Esposito RU. Euphorigenic drugs: effects on the reward pathways of the brain. Federation Proceedings. 1979;38:2473–2476. [PubMed]
  61. Land BB, Bruchas MR, Lemos JC, Xu M, Melief EJ, Chavkin C. The dysphoric component of stress is encoded by activation of the dynorphin kappa-opioid system. Journal of Neuroscience. 2008;28:407–414. [PMC gratis Artikel] [PubMed]
  62. Le Doux JE. Emotion circuits in the brain. Annual Review of Neuroscience. 2000;23:155–184. [PubMed]
  63. Lindholm S, Ploj K, Franck J, Nylander I. D'Repréventioun vun Ethanol huet indirekt kuerz- a laangfristeg Verännerungen vun der Enkephalin- a Dynorphingewënnskonzentraktioun am Réierenhimmel. Alkohol. 2000; 22: 165-171. [PubMed]
  64. Liu J, Schulteis G. Brain reward deficits accompany naloxone-precipitated withdrawal from acute opioid dependence. Pharmacology Biochemistry and Behavior. 2004;79:101–108. [PubMed]
  65. Logan GD, Schachar RJ, Tannock R. Impulsivity and inhibitory control. Psychological Science. 1997;8:60–64.
  66. Mantsch JR, Yuferov V., Mathieu-Kia AM, Ho A, Kreek MJ. Effet vum erweiterte Zougang zu héich Versäuerung vu Kakain-Dosen op d'Selbstverwaltung, Kokainindustrie an d'Gehir vum mRNA-Niveau an Ratten. Psychopharmacologie. 2004; 175: 26-36. [PubMed]
  67. Markou A, Koob GF. Post-cocaine anhedonia: an animal model of cocaine withdrawal. Neuropsychopharmacology. 1991;4:17–26. [PubMed]
  68. Markou A, Koob GF. Intracranial self-stimulation thresholds as a measure of reward. In: Sahgal A, editor. Behavioural Neuroscience: A Practical Approach. Vol. 2. IRL Press; Oxford: 1993. pp. 93–115.
  69. Martin WR, Eades CG. A comparison between acute and chronic physical dependence in the chronic spinal dog. Journal of Pharmacology and Experimental Therapeutics. 1964;146:385–394. [PubMed]
  70. Martinez D, Narendran R, Foltin RW, Slifstein M, Hwang DR, Broft A, Huang Y, Cooper TB, Fischman MW, Kleber HD, Laruelle M. Amphetamine-induced dopamine release: markedly blunted in cocaine dependence and predictive of the choice to self-administer cocaine. American Journal of Psychiatry. 2007;164:622–629. [PubMed]
  71. Martin-Solch C, Magyar S, Kunig G, Missimer J, Schultz W, Leenders KL. Changes in brain activation associated with reward processing in smokers and nonsmokers: a positron emission tomography study. Experimental Brain Research. 2001;139:278–286. [PubMed]
  72. McEwen BS. Allostasis and allostatic load: implications for neuropsychopharmacology. Neuropsychopharmacology. 2000;22:108–124. [PubMed]
  73. Melis M, Spiga S, Diana M. The dopamine hypothesis of drug addiction: hypodopaminergic state. International Review of Neurobiology. 2005;63:101–154. [PubMed]
  74. Merlo-Pich E, Lorang M, Yeganeh M, Rodriguez de Fonseca F, Raber J, Koob GF, Weiss F. Increase of extracellular corticotropin-releasing factor-like immunoreactivity levels in the amygdala of awake rats during restraint stress and ethanol withdrawal as measured by microdialysis. Journal of Neuroscience. 1995;15:5439–5447. [PubMed]
  75. Moeller FG, Barratt ES, Dougherty DM, Schmitz JM, Swann AC. Psychiatric aspects of impulsivity. American Journal of Psychiatry. 2001;158:1783–1793. [PubMed]
  76. Mucha RF, Herz A. Motivational properties of kappa and mu opioid receptor agonists studied with place and taste preference conditioning. Psychopharmacology. 1985;86:274–280. [PubMed]
  77. Nestler EJ. Molecular basis of long-term plasticity underlying addiction. Nature Reviews Neuroscience. 2001;2:119–128. [PubMed]
  78. Nestler EJ. Historical review: molecular and cellular mechanisms of opiate and cocaine addiction. Trends in Pharmacological Sciences. 2004;25:210–218. [PubMed]
  79. Nestler EJ. Is there a common molecular pathway for addiction? Nature Neuroscience. 2005;8:1445–1449. [PubMed]
  80. Nestler EJ, Malenka RC. The addicted brain. Scientific American. 2004;290:78–85. [PubMed]
  81. Neugebauer V, Li W, Vogel GC, Han JS. D'Amyldaala an d'persistente Schief. Neurologen. 2004; 10: 221-234. [PubMed]
  82. Nye HE, Nestler EJ. Induktioun vu chroneschen Fos-bezogenen Antigenen am Rassehous mat chronescher Morphinverabscheidung. Molekulare Pharmakologie. 1996; 49: 636-645. [PubMed]
  83. O’Dell LE, Roberts AJ, Smith RT, Koob GF. Enhanced alcohol self-administration after intermittent versus continuous alcohol vapor exposure. Alcoholism: Clinical and Experimental Research. 2004;28:1676–1682. [PubMed]
  84. Olds J, Milner P. Positiv Verstärkung déi duerch elektresch Stimulatioun vum Septimalgebitt an aner Regiounen vu Rassehänzer produzéiert gëtt. Journal of Vergläichend a Physiologesch Psychologie. 1954; 47: 419-427. [PubMed]
  85. Olive MF, Koenig HN, Nannini MA, Hodge CW. Elevated extracellular CRF levels in the bed nucleus of the stria terminalis during ethanol withdrawal and reduction by subsequent ethanol intake. Pharmacology Biochemistry and Behavior. 2002;72:213–220. [PubMed]
  86. Orsini C, Koob GF, Pulvirenti L. Dopamine partial agonist reverses amphetamine withdrawal in rats. Neuropsychopharmacology. 2001;25:789–792. [PubMed]
  87. Iwwerstreet DH, Knapp DJ, Breese GR. Modulatioun vu méi Ethanol-Récktrouwen-induzéiert Angst-ähnlechen Verhalen vu CRF a CRF1 receptors. Pharmacology Biochemistry and Behavior. 2004;77:405–413. [PMC gratis Artikel] [PubMed]
  88. Paterson NE, Markou A. Increased motivation for self-administered cocaine after escalated cocaine intake. Neuroreport. 2003;14:2229–2232. [PubMed]
  89. Paterson NE, Myers C, Markou A. Effects of repeated withdrawal from continuous amphetamine administration on brain reward function in rats. Psychopharmacology. 2000;152:440–446. [PubMed]
  90. Pfeiffer A, Brantl V., Herz A, Emrich HM. Psychotomimese gëtt duerch κ opiate Rezeptoren vermittelt. Wëssenschaft. 1986; 233: 774-776. [PubMed]
  91. Pulvirenti L, Balducci C, Piercy M, Koob GF. Characterization of the effects of the partial dopamine agonist terguride on cocaine self-administration in the rat. Journal of Pharmacology and Experimental Therapeutics. 1998;286:1231–1238. [PubMed]
  92. Pulvirenti L, Koob GF. Lisuride reduces psychomotor retardation during withdrawal from chronic intravenous amphetamine self-administration in rats. Neuropsychopharmacology. 1993;8:213–218. [PubMed]
  93. Rachlin H, Green L. Commitment, choice and self-control. Journal of the Experimental Analysis of Behavior. 1972;17:15–22. [PMC gratis Artikel] [PubMed]
  94. Rasmussen DD, Boldt BM, Bryant CA, Mitton DR, Larsen SA, Wilkinson CW. Chronic daily ethanol and withdrawal: 1. Long-term changes in the hypothalamo-pituitary-adrenal axis. Alcoholism: Clinical and Experimental Research. 2000;24:1836–1849. [PubMed]
  95. Rassnick S, Heinrichs SC, Britton KT, Koob GF. Microinjection of a corticotropin-releasing factor antagonist into the central nucleus of the amygdala reverses anxiogenic-like effects of ethanol withdrawal. Brain Research. 1993;605:25–32. [PubMed]
  96. Rattan AK, Koo KL, Tejwani GA, Bhargava HN. The effect of morphine tolerance dependence and abstinence on immunoreactive dynorphin (1–13) levels in discrete brain regions, spinal cord, pituitary gland and peripheral tissues of the rat. Brain Research. 1992;584:207–212. [PubMed]
  97. Richter RM, Weiss F. In vivo CRF Verëffentlechung bei Rass Amygdala ass gekuckt beim Kokain-Récktrëtt bei der Selbstverwaltung vu Ratten. Synapse. 1999; 32: 254-261. [PubMed]
  98. Rimondini R, Arlinde C, Sommer W, Heilig M. Long-lasting increase in voluntary ethanol consumption and transcriptional regulation in the rat brain after intermittent exposure to alcohol. FASEB Journal. 2002;16:27–35. [PubMed]
  99. Rivier C, Bruhn T, Vale W. Effect of ethanol on the hypothalamic-pituitary-adrenal axis in the rat: role of corticotropin-releasing factor (CRF) Journal of Pharmacology and Experimental Therapeutics. 1984;229:127–131. [PubMed]
  100. Rodriguez de Fonseca F, Carrera MRA, Navarro M, Koob GF, Weiss F. D'Aktivatioun vum Kortikotropin-Releasefaktor am limbesche System während der Cannabinoidentraff. Wëssenschaft. 1997; 276: 2050-2054. [PubMed]
  101. Rossetti ZL, Hmaidan Y, Gessa GL. Marked inhibition of mesolimbic dopamine release: a common feature of ethanol, morphine, cocaine and amphetamine abstinence in rats. European Journal of Pharmacology. 1992;221:227–234. [PubMed]
  102. Russell MAH. What is dependence? In: Edwards G, editor. Drugs and Drug Dependence. Lexington Books; Lexington MA: 1976. pp. 182–187.
  103. Sanchis-Segura C, Spanagel R. Behavioural assessment of drug reinforcement and addictive features in rodents: an overview. Addiction Biology. 2006;11:2–38. [PubMed]
  104. Sarnyai Z, Biro E, Gardi J, Vecsernyes M, Julesz J, Telegdy G. Brain corticotropin-releasing factor mediates “anxiety-like” behavior induced by cocaine withdrawal in rats. Brain Research. 1995;675:89–97. [PubMed]
  105. Schulteis G, Markou A, Cole M, Koob G. Decreased brain reward produced by ethanol withdrawal. Proceedings of the National Academy of Sciences USA. 1995;92:5880–5884. [PMC gratis Artikel] [PubMed]
  106. Schulteis G, Markou A, Gold LH, Stinus L, Koob GF. Relative sensitivity to naloxone of multiple indices of opiate withdrawal: a quantitative dose-response analysis. Journal of Pharmacology and Experimental Therapeutics. 1994;271:1391–1398. [PubMed]
  107. Schulteis G, Morse AC, Liu J. Repeated experience with naloxone facilitates acute morphine withdrawal: potential role for conditioning processes in acute opioid dependence. Pharmacology Biochemistry and Behavior. 2003;76:493–503. [PubMed]
  108. Schulteis G, Yackey M, Risbrough V, Koob GF. Anxiogenic-like effects of spontaneous and naloxone-precipitated opiate withdrawal in the elevated plus-maze. Pharmacology Biochemistry and Behavior. 1998;60:727–731. [PubMed]
  109. Self DW, McClenahan AW, Beitner-Johnson D, Terwilliger RZ, Nestler EJ. Biochemical adaptations in the mesolimbic dopamine system in response to heroin self-administration. Synapse. 1995;21:312–318. [PubMed]
  110. Shaw-Lutchman TZ, Barrot M, Wallace T, Gilden L, Zachariou V, Impey S, Duman RS, Storm D, Nestler EJ. Regional and cellular mapping of cAMP response element-mediated transcription during naltrexone-precipitated morphine withdrawal. Journal of Neuroscience. 2002;22:3663–3672. [PubMed]
  111. Shippenberg TS, Koob GF. Recent advances in animal models of drug addiction and alcoholism. In: Davis KL, Charney D, Coyle JT, Nemeroff C, editors. Neuropsychopharmacology: The Fifth Generation of Progress. Lippincott Williams and Wilkins; Philadelphia: 2002. pp. 1381–1397. 2002.
  112. Slawecki CJ, Thorsell AK, Khoury AE, Mathe AA, Ehlers CL. Increased CRF-like and NPY-like immunoreactivity in adult rats exposed to nicotine during adolescence: relation to anxiety-like and depressive-like behavior. Neuropeptides. 2005;39:369–377. [PubMed]
  113. Solomon RL, Corbit JD. An opponent-process theory of motivation: 1. Temporal dynamics of affect. Psychological Review. 1974;81:119–145. [PubMed]
  114. Song ZH, Takemori AE. Stimulation by corticotropin-releasing factor of the release of immunoreactive dynorphin A from mouse spinal cords in vitro. European Journal of Pharmacology. 1992;222:27–32. [PubMed]
  115. Spangler R, Unterwald EM, Kreek MJ. Binge” cocaine administration induces a sustained increase of prodynorphin mRNA in rat caudate-putamen. Molecular Brain Research. 1993;19:323–327. [PubMed]
  116. Specio SE, Wee S, O'Dell LE, Boutrel B, Zorrilla EP, Koob GF. CRF1 Rezeptor Antagonisten attenuéieren eskaléiert Kokain Selbstverwaltung bei Ratten. Psychopharmakologie. 2008;196:473-482. [PMC gratis Artikel] [PubMed]
  117. Sterling P, Eyer J. Allostasis: a new paradigm to explain arousal pathology. In: Fisher S, Reason J, editors. Handbook of Life Stress, Cognition and Health. John Wiley; Chichester: 1988. pp. 629–649.
  118. Stinus L, Cador M, Zorrilla EP, Koob GF. Buprenorphine and a CRF1 antagonist block the acquisition of opiate withdrawal-induced conditioned place aversion in rats. Neuropsychopharmacology. 2005;30:90–98. [PubMed]
  119. Stinus L, Caille S, Koob GF. Opiate withdrawal-induced place aversion lasts for up to 16 weeks. Psychopharmacology. 2000;149:115–120. [PubMed]
  120. Stinus L, Le Moal M, Koob GF. Nucleus accumbens and amygdala are possible substrates for the aversive stimulus effects of opiate withdrawal. Neuroscience. 1990;37:767–773. [PubMed]
  121. Thorsell A, Rapunte-Canonigo V, O’Dell L, Chen SA, King A, Lekic D, Koob GF, Sanna PP. Viral vector-induced amygdala NPY overexpression reverses increased alcohol intake caused by repeated deprivations in Wistar rats. Brain. 2007;130:1330–1337. [PMC gratis Artikel] [PubMed]
  122. Thorsell A, Slawecki CJ, Ehlers CL. Effects of neuropeptide Y and corticotropin-releasing factor on ethanol intake in Wistar rats: interaction with chronic ethanol exposure. Behavioural Brain Research. 2005a;161:133–140. [PubMed]
  123. Thorsell A, Slawecki CJ, Ehlers CL. Effects of neuropeptide Y on appetitive and consummatory behaviors associated with alcohol drinking in wistar rats with a history of ethanol exposure. Alcoholism: Clinical and Experimental Research. 2005b;29:584–590. [PubMed]
  124. Tucci S, Cheeta S, Seth P, Datei SE. Kortikotropin freet Faktorantagonist, α-helical CRF9-41, reverses nicotine-induced conditioned, but not unconditioned, anxiety. Psychopharmacology. 2003;167:251–256. [PubMed]
  125. Valdez GR, Sabino V, Koob GF. Increased anxiety-like behavior and ethanol self-administration in dependent rats: reversal via corticotropin-releasing factor-2 receptor activation. Alcoholism: Clinical and Experimental Research. 2004;28:865–872. [PubMed]
  126. Valdez GR, Zorrilla EP, Roberts AJ, Koob GF. De Antagonismus vum Corticotropin-Releasereffekt huet d'verstäerkt Reaktiounsfäegkeet zu Stress beobachtet, déi beobachtet ginn während enger laangwiereger Ethanol-Abstinenz. Alkohol. 2003; 29: 55-60. [PubMed]
  127. Valentino RJ, Foote SL, Page ME. The locus coeruleus as a site for integrating corticotropin-releasing factor and noradrenergic mediation of stress responses. In: Tache Y, Rivier C, editors. Corticotropin-Releasing Factor and Cytokines: Role in the Stress Response. New York Academy of Sciences; New York: 1993. pp. 173–188. (series title: Annals of the New York Academy of Sciences, vol 697) [PubMed]
  128. Valentino RJ, Page ME, Curtis AL. Activation of noradrenergic locus coeruleus neurons by hemodynamic stress is due to local release of corticotropin-releasing factor. Brain Research. 1991;555:25–34. [PubMed]
  129. Van Bockstaele EJ, Colago EE, Valentino RJ. Amygdaloid corticotropin-releasing factor targets locus coeruleus dendrites: substrate for the co-ordination of emotional and cognitive limbs of the stress response. Journal of Neuroendocrinology. 1998;10:743–757. [PubMed]
  130. Van Dyke C, Byck R. Cocaine. Scientific American. 1982;246:128–141. [PubMed]
  131. Vezina P. Sensibiliséierung vun Midbrain-Dopamin-Neuron-Reaktivitéit an der Selbstverwaltung vun Psychomotor-stimulant Medikamenter. Neurologie an Biobehavioraler Rezensiounen. 2004; 27: 827-839. [PubMed]
  132. Volkow ND, Fowler JS. Addiction, a disease of compulsion and drive: involvement of the orbitofrontal cortex. Cerebral Cortex. 2000;10:318–325. [PubMed]
  133. Volkow ND, Fowler JS, Wang GJ. Role of dopamine in drug reinforcement and addiction in humans: results from imaging studies. Behavioural Pharmacology. 2002;13:355–366. [PubMed]
  134. De Volkow ND, Wang GJ, Fowler JS, Logan J, Gatley SJ, Hitzemann R, Chen AD, Dewey SL, Pappas N. Verréngert striatal dopaminergic Responsabilitéit an dégoxifizéierter Kokain-Abhängefähegkeet. Natur. 1997; 386: 830-833. [PubMed]
  135. Walker BM, Koob GF. Pharmakologesch Beweiser fir eng motivational Roll vun κ-Opioid Systemer an Ethanol Ofhängegkeet. Neuropsychopharmacologie. 2008; 33: 643-652. [PMC gratis Artikel] [PubMed]
  136. Walker BM, Rasmussen DD, Raskind MA, Koob GF. α1-Noradrenergic receptor antagonism blocks dependence-induced increases in responding for ethanol. Alcohol. 2008;42:91–97. [PMC gratis Artikel] [PubMed]
  137. Wee S, Mandyam CD, Lekic DM, Koob GF. α1-Noradrenergic system role in increased motivation for cocaine intake in rats with prolonged access. European Neuropsychopharmacology. 2008;18:303–311. [PMC gratis Artikel] [PubMed]
  138. Wee S, Wang Z, Woolverton WL, Pulvirenti L, Koob GF. Effekt vun Aripiprazol, eng partiell D2 receptor agonist, on increased rate of methamphetamine self-administration in rats with prolonged access. Neuropsychopharmacology. 2007;32:2238–2247. [PMC gratis Artikel] [PubMed]
  139. Weiss F, Markou A, Lorang MT, Koob GF. Basal extracellular dopamine levels in the nucleus accumbens are decreased during cocaine withdrawal after unlimited-access self-administration. Brain Research. 1992;593:314–318. [PubMed]
  140. Weiss F, Ciccocioppo R, Parsons LH, Katner S, Liu X, Zorrilla EP, Valdez GR, Ben-Shahar O, Angeletti S, Richter RR. Compulsive drug-seeking behavior and relapse: neuroadaptation, stress, and conditioning factors. In: Quinones-Jenab V, editor. The Biological Basis of Cocaine Addiction. New York Academy of Sciences; New York: 2001. pp. 1–26. (series title: Annals of the New York Academy of Sciences, vol 937) [PubMed]
  141. Weiss F, Parsons LH, Schulteis G, Hyytia P, Lorang MT, Bloom FE, Koob GF. Ethanol self-administration restores withdrawal-associated deficiencies in accumbal dopamine and 5-hydroxytryptamine release in dependent rats. Journal of Neuroscience. 1996;16:3474–3485. [PubMed]
  142. Wikler A. A psychodynamic study of a patient during experimental self-regulated re-addiction to morphine. Psychiatric Quarterly. 1952;26:270–293. [PubMed]
  143. Zhang Z, Schulteis G. Withdrawal from acute morphine dependence is accompanied by increased anxiety-like behavior in the elevated plus maze. Pharmacology Biochemistry and Behavior. 2008;89:392–403. [PMC gratis Artikel] [PubMed]