You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/11267539

Nicotine addiction: Insights from recent animal studies

Article  in  Psychopharmacology · August 2002


DOI: 10.1007/s00213-002-1096-0 · Source: PubMed

CITATIONS READS

61 627

4 authors, including:

Anne-Marie Mathieu-Kia Scott Kellogg


Sanofi Transformational Chairwork Psychotherapy Project
21 PUBLICATIONS   723 CITATIONS    34 PUBLICATIONS   1,809 CITATIONS   

SEE PROFILE SEE PROFILE

Mary Jeanne Kreek


The Rockefeller University
619 PUBLICATIONS   29,339 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Chairwork Psychotherapy View project

Opioid physiology View project

All content following this page was uploaded by Scott Kellogg on 31 May 2014.

The user has requested enhancement of the downloaded file.


Psychopharmacology (2002) 162:102–118
DOI 10.1007/s00213-002-1096-0

REVIEW

A. M. Mathieu-Kia · S. H. Kellogg · E. R. Butelman


M. J. Kreek

Nicotine addiction: insights from recent animal studies

Received: 13 August 2001 / Accepted: 7 March 2002 / Published online: 15 May 2002
© Springer-Verlag 2002

Abstract Rationale: Recent preclinical behavioral and Introduction


neurobiological research has characterized important
behavioral features and has identified neurobiological Over the past 20 years preclinical research studies have
substrates that may underlie nicotine reinforcement and been accumulating on the neurobiology of nicotine
addiction. Objective: To examine recent advances on addiction. This review emphasizes the recent literature
nicotine exposure in preclinical models, from three per- on the pharmacokinetics of nicotine, on animal models
spectives: (a) the chronopharmacokinetics of nicotine, to explore the distinct phases of nicotine addiction, and
(b) behavioral studies on nicotine reinforcement, with- on nicotinic neuropharmacology.
drawal, and reinstatement/relapse, and (c) effects of
nicotine on neurobiological substrates after repeated
exposure. Results: Preclinical studies can be used to Chronopharmacokinetic characteristics of nicotine
operationally model selected aspects of nicotine rein-
forcement, withdrawal, and reinstatement or relapse. Inasmuch as nicotine is thought to be the primary com-
These may be used to investigate the functional in vivo pound in tobacco smoke that establishes and maintains
consequences of acute and long-term changes in neuro- tobacco dependence, knowledge of its pharmacokinetic
nal acetylcholine receptor populations that follow nico- characteristics is important for understanding the daily
tine exposure. Behavioral studies focusing on distinct smoking cycle. Smoking with frequent or deep inhala-
stages of nicotine exposure (e.g., active reinforcement tion is the route of administration that allows for the
vs. cessation or reinstatement) may also be used in parallel most rapid delivery of nicotine to the brain. Nicotine,
with studies on dopaminergic function, a proposed sub- mainly through pulmonary absorption and via arterial
strate for the reinforcing effects of nicotine, and of opioid blood, reaches the brain within a few seconds. This
receptor function, a possible site of neuroadaptations pattern is consistent with drugs that have a high abuse
secondary to nicotine exposure. Conclusions: While no potential (Quinn et al. 1997). As it is distributed to other
single current animal model may capture the experience body tissues, circulating nicotine is rapidly oxidized into
of human smoking or nicotine addiction, increasingly, cotinine, a major metabolite eliminated in urine. In
separate animal models are capturing the full spectrum human smokers, as was also shown in recent data using
of behavioral and neurobiological dimensions of this the stable isotopic nicotine technique (Benowitz et al.
complex condition. 1999), the initial distribution half-life (t1/2α) of nicotine in
the blood is estimated to be 4–10 min, and the elimination
Keywords Preclinical model · Drug-seeking behavior · half-life (t1/2β) is approximately 120 min (Benowitz
Nicotinic · Dopamine · Opioid 1988; Lunell et al. 2000; Perez-Stable et al. 1998). Arterial
nicotine concentrations peak approximately 20 s after
each puff, yielding increases in nicotine concentrations
in the order of 10 ng/ml (Rose et al. 1999). This rapid
arterial nicotine increase following each puff may be
A.M. Mathieu-Kia · S.H. Kellogg (✉) · E.R. Butelman especially relevant to the comparative reinforcing prop-
M.J. Kreek erties of cigarette smoking vs. other forms of nicotine
Rockefeller University, delivery, such as gum, transdermal patch, and nasal
Laboratory on the Biology of Addictive Diseases, Box 171,
1230 York Avenue, New York, NY 10021-6399, USA spray (e.g., Gourlay and Benowitz 1997; Perkins et al.
e-mail: kellogs@mail.rockefeller.edu 1994; Schneider et al. 1996). After termination of each
Tel.: +1-212-3278247, Fax: +1-212-3278574 cigarette, the blood level of nicotine rapidly peaks and
103

Fig. 1 Proposed circadian profile of circulating nicotine in regular tine is slightly longer than in the blood (Plowchalk et al.
smokers. This theoretical schematic drawing models the intermittent 1992; Rowell and Li 1997; Sastry et al. 1995). As the
peaking-falling pattern during daylight hours and the circadian
fluctuations in nicotine concentration in the blood of human tobacco number of injections increases, the brain levels of nico-
smokers. Superimposed on these two different patterns of oscillation, tine, which are on average two or three times higher than
the overall level of nicotine gradually increases across the morning, those in the blood (Donny et al. 2000; Rowell and Li
reaches a plateau during the afternoon to the late evening, and 1997), increase and reach a plateau which is difficult to
finally decreases during the night to achieve low concentrations by
the next morning. Ideally, in animal models of nicotine addiction,
extrapolate from the blood values (Ghosheh et al. 2001).
we should be able to reproduce a similar profile with respect to the Additional work is needed to establish the pharmaco-
analogous circadian rhythm. White horizontal bars Daytime hours; kinetic profile of nicotine in specific brain regions and in
black horizontal bars nighttime hours. This drawing is arithmetically studies of behavioral models of addiction. Furthermore,
constructed and is based on the short and long metabolic half-lives recent experiments have contributed to the development
of nicotine using a model of 8 cigarette smoking episodes a day.
This graph is inspired by, and very close to, the experimental data of possible novel agonist pharmacotherapies for breaking
obtained in human smokers (Benowitz et al. 1982) the cycle of nicotine addiction in humans. For example,
recent studies have determined that nornicotine (N-de-
methylated nicotine; an alkaloid present in tobacco, and
falls. In addition to this intermittent spike pattern, the
also a minor nicotine metabolite), was weakly self-
overall level of nicotine across the daylight period grad-
administered by rats but could antagonize intravenous
ually increases in the morning, reaches maximal level
nicotine self-administration (SA; Bardo et al. 1999;
during the afternoon and late evening, and finally falls
Green et al. 2000).
during the night to achieve low but potentially active
concentrations by the next morning (Benowitz et al.
1982). This pattern repeats itself daily (Fig. 1). The Acquisition and maintenance of nicotine-seeking
potential of nicotine to desensitize nAChRs should also behavior
be considered in this context, since receptor desensitiza-
tion may occur in the presence of relatively low nicotine Numerous experiments have shown that nicotine serves
concentrations (such as may be encountered during as a positive reinforcer, and this has been extensively
“trough” periods; e.g., Marks et al. 1996; see also documented in diverse animal species (almost exclusively
“Primary molecular site of action: the neuronal nicotinic male) including primates, dogs, and rodents (for reviews
acetylcholine receptors,” below). see Goldberg et al. 1983; Henningfield and Goldberg
In rodent models nicotine is also quickly absorbed 1983; Perkins et al. 1999; Stolerman 1999). Although the
and its blood concentration rapidly declines in a biexpo- conditioned place preference assay (Calcagnetti and
nential fashion. In rats after ten puffs of cigarette smoke, Schecter 1994; Clark and Fibiger 1987; Jorenby et al.
the t1/2α and t1/2β values of nicotine are 5 and 86 min, 1990; Risinger and Oakes 1995) has been used in some
respectively (Rotenberg et al. 1980). Similarly, after an studies, the intravenous SA assay has been favored as a
acute intra-arterial or intravenous administration of nico- more direct way of demonstrating that nicotine initiates
tine mean t1/2α and t1/2β of nicotine are approximately 10 and sustains drug-seeking behavior (self-administration
and 90 min, respectively (Adir et al. 1976; Kyerematen by other routes has also been studied, e.g., Glick et al.
et al. 1988; Miller et al. 1977; Plowchalk et al. 1992; 1996). However, to perfectly reproduce the route,
Sastry et al. 1995). As for acute administration, calcula- frequency, chronology, and intake pattern of smoking the
tions of the pharmacokinetic parameters in blood from ideal approach would be to have an animal actually
rats chronically treated with nicotine yield a global t1/2 of smoke cigarettes or self-inhale smoke on a smoking
44–49 min. This is a consistent value regardless of machine. In fact, rhesus monkeys and baboons do smoke
whether rats receive the subcutaneous nicotine injections after this behavior has been established in an operant
two, four, or eight times a day over a 10-day period situation. In some extreme cases they would smoke up to
(Rowell and Li 1997). In the brain the global t1/2 of nico- 40 cigarettes/day, for periods lasting up to 3 years (Ando
104

and Yanagita 1981; Rogers et al. 1985, 1988). Despite access; nevertheless, the rats rapidly initiated SA behav-
the clear analogy to human behavior, studies using ior. However, there was a relatively low rate of SA
smoking protocols with nonhuman primates have been behavior, and there was not a general distinction between
rare in the last decade, probably due to material and responding on the active and inactive levers. In these
technical demands. conditions of unlimited drug access the maximal nicotine
As is commonly observed in intravenous SA assays, a intake was approximately 0.2–1.4 mg/kg per day, vs.
crucial experimental parameter that governs the acquisi- 0.4–0.9 mg/kg per hour in the restricted access approach
tion of SA behavior is the unit dose of nicotine (all doses (Corrigall and Coen 1989; Cox et al. 1984; Donny et al.
in text are presented as nicotine base, unless otherwise 1995).
stated). The optimal acquisition dose is thought to be Other important factors (both similar to and different
around 0.01–0.03 mg/kg per intravenous injection in from the human smoking cycle) that could influence the
many animal species (Corrigall and Coen 1989; Cox et nicotine-seeking behavior also include previous experi-
al. 1984; Donny et al. 1995, 1998; Goldberg et al. 1981; mental administration or SA of various compounds such
Henningfield and Goldberg 1983; Risner and Goldberg as cocaine (Epping-Jordan et al. 1999; Tessari et al.
1983; Shoaib and Stolerman 1999; Shoaib et al. 1997; 1995), caffeine (Tanda and Goldberg 2000) or nicotine
Suto et al. 2001; Tessari et al. 1995). The maintenance of itself. Thus, noncontingent exposure to nicotine can
nicotine-seeking behavior in animal models is also dose facilitate the later acquisition of nicotine SA (Shoaib et
dependent, and an inverted “U” function is usually al. 1997), the development of conditioned place preference
found. However, the overall function may be shallower (Shoaib et al. 1994), and the increased intake of nicotine
for nicotine than for other reinforcing drugs, and may as shown in a two-bottle choice study (Maehler et al.
suggest comparatively less dose regulation for nicotine 2000). Although the direct parallel between these findings
vs. other reinforcing drugs (Lynch and Carroll 1999; and human smoking behavior is unclear, these findings
Rose and Corrigall 1997). The inverted “U” shape dose- may reflect the process by which a relatively small
effect curve has also been observed in conditioned place number of cigarette exposures puts an individual at high
preference assays with mice (e.g., Risinger and Oakes risk for developing nicotine addiction.
1995) and rats (Fudala and Iwamoto 1986). Within the SA literature, there has been a growing
Nicotine-seeking behavior has been examined in a interest in the difference between the fixed-ratio and the
variety of procedural conditions, both analogous to and progressive-ratio schedules. Markou et al. (1993) have
different from the human smoking cycle. The issue of proposed that the fixed-ratio schedules are a measure of
restricted versus unlimited access to nicotine is central the “hedonic” value of the drug while the progressive-
because of concerns about how to best mimic human ratio schedules are analogous to the experience of moti-
smoking behavior with respect to both the frequency of vational incentive or “drug craving.” The “breaking
cigarette use and the sleep-wake cycle. Restricted access point” is seen as a potential measure of the “cost” of the
to nicotine has been a common model, which may not drug. Donny et al. (2000) used this model to evaluate
capture the experience of human smokers who have sex-based, motivational-incentive differences in rats
access to tobacco at any time. In light of recent legislative that could potentially parallel the greater difficulty that
interventions in the United States, access to tobacco is women have with the cessation of smoking (Perkins et
becoming more restricted, and increasing numbers of al. 1999).
smokers may actually be following a fixed or variable It has been recently proposed that stimuli associated
interval pattern of SA (i.e., possibly similar to a “binge” with smoking or with nicotine SA procedures (e.g., stim-
use pattern). As investigations go forward, it will be ulus lights) acquire conditioned reinforcer status, and
increasingly useful to systematically compare restricted that this contributes to the maintenance of tobacco or
access, unlimited access, and “binge” models. nicotine SA (Balfour et al. 2000; Di Chiara 2000; Donny
In an example of research utilizing schedules of repet- et al. 1999). This may be analogous to the sensory stimuli
itive nicotine exposure, nicotine was administered ad (e.g., in the respiratory tract) that may be temporally
libitum to mice in their sole source of drinking water. associated with rapid arterial nicotine increases during
Mice, as nocturnal animals, drink primarily during their cigarette smoking in humans (e.g., Rose et al. 1999).
active phase, and as a result their plasma levels of nico-
tine follow a circadian high-to-low profile similar to that
reported in human smokers (for reviews see Pietila and Extinction of nicotine-seeking behavior
Ahtee 2000). This oral intake approach leads to a rela- and withdrawal
tively slow nicotine absorption pattern that may not fully
capture common cigarette use patterns. Valentine et al. Smoking cessation or periods of forced nicotine absti-
(1997), in an attempt to more closely approximate nence have been examined in animals undergoing
human nicotine use, developed a model in which rats extinction of nicotine-seeking behavior or cessation after
were allowed unlimited access to nicotine using an intra- chronic experimenter-delivered administration. In rodent
venous SA assay while respecting the circadian/meal SA experiments studying extinction, when saline is sub-
cycles. The four doses used (0.00375–0.03 mg/kg) were stituted for nicotine, an overall decline in drug-seeking
low compared to the commonly used doses in restricted behavior is usually observed. During the first substitu-
105

tion session some rats begin pressing the bar less fre- show this same change in the ICSS (Epping-Jordan et al.
quently while others demonstrated a typical “extinction 1998). Nonetheless, once the continuous nicotine infu-
burst” – a rapid but short-lived increase in bar pressing sion is terminated, the spontaneous withdrawal condition
levels (Corrigall and Coen 1989; Donny et al. 1995; elevates ICSS thresholds. The rat may be postulated to
Shoaib et al. 1999). Interestingly, the drug-seeking be- be in a “dysphoric” state, a condition that may parallel
havior remains above control levels for 4–13 sessions the “negative mood state” seen in smokers, which is
before being completely extinguished (Chiamulera et al. typified by irritability, lack of concentration, anxiety, and
1996; Donny et al. 1995; Tessari et al. 1995); this resis- craving (Benowitz 1988; see also below). Similarly, con-
tance to extinction further supports the view of nicotine ditioned place aversion induced by precipitated nicotine
as a powerfully addictive compound (e.g., DiChiara withdrawal may provide a sensitive measure of internal
2000). states (e.g., discriminative or hedonic) produced by nico-
Early and long-lasting behavioral signs of extinction tine withdrawal in animals (Ise et al. 2000; Watkins et al.
were also produced by preadministration with a nicotinic 2000a).
antagonist (Goldberg and Spealman 1982; Shoaib et al. Other studies have also attempted to examine the
1997). “subjective” or interoceptive state of animals when they
Nicotine withdrawal in animals consists of a syn- experience nicotine withdrawal. Spontaneous withdrawal
drome of spontaneous or precipitated physical signs – in rats has been marked by an increased “sensorimotor
similar to those found during opiate withdrawal (including reactivity” response to acoustic startle (Acri et al. 1991;
teeth-chattering/chews, yawns, abdominal writhes/gasps, Harris et al. 1986; Helton et al. 1993, 1997; Rasmussen
ptosis, wet shakes/tremors, rearing, and jumping). In- et al. 1997).
creases in body weight, decreases in spontaneous activity, The withdrawal state after chronic, repeated nicotine
hyperphagia, and hyperdipsia are also used to evaluate administration is also notable for its significant disrup-
the withdrawal intensity (Hildebrand et al. 1997; Levin tion of operant behaviors not directly linked to the drug-
et al. 1987; Malin et al. 1992). In rats the discontinuation seeking behavior. When nicotine access is terminated,
of subcutaneous nicotine infusions (3.2–3.6 mg/kg per the reinforcement potential of food or sweetened solu-
day) produces a constellation of spontaneous behavioral tion is suppressed, and this interference is immediately
signs during the 2 days following cessation (Hildebrand reversed when nicotine is again available (Carroll et al.
et al. 1997; Malin et al. 1992, 1994; Watkins et al. 1989; Corrigall and Coen 1989). Similarly, the abrupt
2000a). Nicotinic antagonists are also able to precipitate cessation of nicotine administration in rats trained on a
all of these abstinence symptoms, and a subsequent nico- shock-avoidance schedule is also associated with a re-
tine injection (e.g., 0.15–0.4 mg/kg subcutaneously) duction in lever-pressing responses and with an increase
attenuates their intensity (Hildebrand et al. 1997; Malin in the number of shocks received (for a review see
et al. 1994, 1998). In a study using daily intermittent Balfour 1991). These behavioral alterations are seen as
nicotine injections in mice, mild somatic withdrawal acute signs of withdrawal and could be compared to
signs were reported between 24 and 48 h after the final the loss of the “psychological benefits” of smoking
bolus, and some of them persisted for 3–4 days (Isola et commonly described by smokers during abstinence.
al. 1999). When these studies are contrasted, continuous
and intermittent exposures share similar withdrawal
symptom while following different temporal courses. It Reinstatement model of nicotine-seeking behavior
is to be noted that none of the studies to date on with- and relapse
drawal signs have used a SA model (to our knowledge),
and this approach would more closely mimic the behavior Experimental relapse from nicotine addiction has been
of smokers. examined through the reinstatement of previously extin-
The intracranial self-stimulation (ICSS) model has guished nicotine-seeking behaviors, using spontaneous
also been able to measure the reinforcing properties of recovery conditions, nicotine priming doses, or stress-
nicotine exposure and the state change (or “presumed inducing stimuli. Animals removed from their operant
dysphoria”) that follows its termination. In the ICSS chamber or in saline substitution protocols spontaneously
model the effect of nicotine is thought to mimic the power recover drug-seeking behavior at levels comparable to
and rewarding effect of each electrical stimulation (for a those observed during the preextinction period within
review see Wise 1996). As a result the ICSS threshold is one session (Donny et al. 1995; Goldberg et al. 1981;
lower for the nicotine-treated animals than for those Valentine et al. 1997). Similarly, after 6–20 sessions of
receiving vehicle. This stimulus substitution is reached saline substitution, the availability of nicotine progres-
as early as the second day of administration and lasts as sively reinitiates SA to preextinction levels (Goldberg et
long as nicotine is repeatedly injected and not blocked al. 1981; Tessari et al. 1995). Lastly, after a 3-week period
with nicotinic antagonists (Bauco and Wise 1994; of nicotine withdrawal, rats reexposed to the drug-taking
Bespalov et al. 1999; Bozarth et al. 1998a, 1998b; environment spontaneously recover their previous behavior
Huston-Lyons et al. 1993; Ivanová and Greenshaw (Shaham et al. 1997). Whatever the experimental condi-
1997). In a contrasting effect, animals maintained on a tion, spontaneous recovery of SA occurs after either a
continuous infusion of nicotine by osmotic pump do not short or long cessation period, and this phenomenon
106

illuminates the connection between environmental factors anxiogenic effects of nicotine and nicotine withdrawal.
and relapse. Ouagazzal et al. (1999) found that nicotine had no effect
Another approach to study relapse is to use an acute at the lower doses, but was anxiogenic in the elevated
priming nicotine injection to reinstate nicotine SA. An plus-maze test at higher doses. Irvine et al. (2001a) in a
acute priming nicotine dose (either intravenous or subcu- study with male Wistar rats examined the effects of nico-
taneous) reinstates nicotine SA after 4–13 days of saline tine in the social interaction test after 4 weeks of nicotine
substitution (Chiamulera et al. 1996; Shaham et al. SA and then, again, 24 to 72 h after its cessation. Rats
1997). Interestingly, higher levels of reinstatement of that self-administered nicotine showed less interaction
nicotine SA are observed with a lower priming dose of with other rats in the social interaction test than the
nicotine (Chiamulera et al. 1996). Thus, in addition to control group (an apparent nicotine-induced, anxiety-like
potentially conditioned environmental stimuli, nicotine- effect). These group differences disappeared during the
seeking behavior can also be reinitiated through noncon- nicotine cessation period. Irvine et al. (2001a) concluded
tingent nicotine reexposure. that “nicotine self administration is not maintained
Intermittent footshock as a stressor is also an effective because of its anxiolytic effect, but despite or because
method of reinstating nicotine-seeking behavior – even of, its anxiogenic effect.” They also pointed out that
after periods ranging from 5 days to 2 weeks (e.g., Buczek the findings of the animal studies were in conflict
et al. 1999). This supports the clinical observation that with common assumptions about the behavior of human
stress can be a major precipitant of relapse; continued smokers.
work in this area is crucial as this is a major therapeutic The cessation of nicotine administration also appears
intervention point for smoking-cessation medications. to induce anxiety in animals, and this has been proposed
as paralleling the withdrawal symptoms that serve as
negative reinforcers working to maintain the use of
Animal models of the “psychological” experience tobacco. Classical rodent models of “anxiety,” such as
of smoking the black/white box test (Costall et al. 1989), the elevated
plus-maze test (Irvine et al. 2001b), and the social inter-
A number of recent studies have attempted to model the action test (Irvine et al. 1999), have been used to evaluate
wide variety of experiences and use patterns that smokers the nicotine withdrawal state. An anxiogenic response
report. Areas of investigation have included psychological first appeared within 8–96 h following the cessation of
mood states, such as depression and anxiety, and the chronic experimenter-delivered nicotine in these settings.
cooccurring use of other substances, such as caffeine and However, anxiogenic effects were not detected 1 or
alcohol. 3 days following the end of a 4-week period of intrave-
The relationship between nicotine and depression has nous nicotine (0.03 mg/kg per injection) SA (Irvine et al.
been one of growing interest. Studies have shown that 2001a; see above). To summarize, the overall relation-
cigarette smoking is quite common among depressed ship between nicotine and anxiety remains complex and
individuals and that individuals with a long history of unclear.
depression are less likely to succeed at smoking cessation Tobacco use and cigarette smoking often occur con-
(Tizabi et al. 1999). In fact, buproprion, an antidepres- currently with the use of other substances. Recent studies
sant, is prescribed to patients seeking to discontinue have attempted to examine two common combinations –
smoking (NIDA 1998). nicotine and caffeine and nicotine and alcohol. Shoaib et
Semba et al. (1998) used a “learned helplessness” al. (1999) reported that caffeine markedly increased the
model that involved shock and an escape test in male rats reinforcing properties of nicotine using an SA model,
and found that chronic administration of nicotine may and its consequent removal diminished the reinforcing
have an apparent antidepressant effect, and that meca- properties of nicotine. This finding supports a role for
mylamine, a nicotinic antagonist, reversed this effect of caffeine as a possible contributor to nicotine SA and
nicotine. Similarly, Tizabi et al. (1999) examined the relapse (Shoaib et al. 1999).
apparent antidepressant effects of nicotine in the “forced Other studies examined the interaction between nico-
swim” model of depression with two related rat lines. tine and alcohol. Here the emphasis has been on the
The Flinders Sensitive Line (FSL) has been specifically effect of nicotine on alcohol consumption. Nadel and
bred to be a “depression-prone” rat, and the Flinders Samson (1999), using male Long-Evans rats, found that
Resistant Line (FRL) has been bred to be a “depression- nicotine had only a slight effect on increasing alcohol
resistant” rat. After both acute and chronic (14-day) SA. Lê et al. (2000), using male Wistar rats, found a
administration of nicotine, immobility was decreased in complex relationship between exposure to nicotine and
the “depressed” FSL rat but not in the FRL rats. Djurič et alcohol consumption. Nicotine only impacted alcohol
al. (1999), also using the forced swim-test with female SA at higher doses (0.8 mg/kg), and the effect was
FSL and FRL rats (found after 14 days of oral nicotine biphasic, as it actually suppressed alcohol consumption
ingestion), an antidepressant effect of nicotine that did at first and then increased it. The authors suggested that
not interact with genetic background. this biphasic effect of nicotine on alcohol consumption
Another area of investigation is the relationship be- might explain conflicting reports about the relationship
tween smoking and anxiety, focusing on the anxiolytic/ between these substances in the literature.
107

Gender and genetics as mediating variables


in the effects of nicotine
In a comprehensive review, Perkins et al. (1999) dis-
cussed the emerging evidence that men and women differ
in their responses to smoking in general. The human data
point to a pattern in which the relative importance of
“nonnicotine” factors that are involved in smoking
behavior is greater for women than for men. This may
include such factors as pleasure from the smoke itself,
social reinforcements for smoking, comfort from having
something to manipulate in social situations, and gratifi-
cation from oral stimulation. Faraday et al. (1999) have
also pointed out that women are more inclined to see
smoking as a useful mechanism for coping with social Fig. 2 Anatomical sites of action for nicotine in the rat brain. This
situations or unpleasant affective states. Perkins et al. sagittal section illustrates the distribution of high-affinity binding
sites for [3H]nicotine. SC Superior colliculus, AV anteroventral
(1999) went on to point out that these differences have nucleus, VL ventrolateral nucleus, Rt reticular nucleus, Cpu dorsal
had troubling consequences for women smokers who are striatum, Acb nucleus accumbens, SNC/VTA substantia nigra com-
seeking to quit; mainly that women have been less suc- pacta/ventral tegmental area. (Slightly modified from Clarke et al.
cessful than men in studies using nicotine-replacement 1985, copyright by the Society for Neuroscience)
interventions. Interestingly, women respond better than
men in studies in which such nonnicotine medications as
variable. Both strain and age differences were found, with
clonidine or buproprion are used (Perkins et al. 1999).
the WKR consuming more ethanol and nicotine than the
The animal literature has supported this pattern to
BNR, and the older BNR consumed less nicotine and
some degree as well. Perkins et al. (1999) note that nico-
ethanol than the younger (Todte et al. 2001). These and
tine appears to be equally reinforcing for males and fe-
other (Robinson et al. 1996; Stolerman et al. 1999) recent
males, but that aspects of the experimental environment
studies therefore support future investigation of the influ-
that may become associated with nicotine may contribute
ence of genetic factors in nicotine SA.
a more significant role in the maintenance of nicotine SA
for females than for males.
Faraday et al. (1999) examined the interaction of gender Primary molecular site of action:
and housing (single versus group) on nicotine-induced the neuronal nicotinic acetylcholine receptors
locomotion in Long-Evans rats in a Plexiglass arena. The
emerging patterns showed that there was an interaction Nicotine specifically binds to the nicotinic receptors
between gender and housing, and nicotine affected the (nAChRs), which mediate initially its effects, and are
arousal and exploration behavior of males, while it widely distributed in the brain of many species including
appeared to have had an “anxiolytic-like” effect on the humans (for a review see Paterson and Nordberg 2000).
female rats. A precise map of the distribution of nAChR (Fig. 2) was
Donny et al. (2000), examined sex differences in SA first established using quantitative autoradiography with
using Sprague-Dawley rats. In this study they used both [3H]nicotine (Clarke et al. 1984, 1985), and later with
fixed-ratio schedules, which they postulated reflect the other approaches such as immunohistochemistry (Deutch
hedonic aspect of nicotine, and progressive-ratio sched- et al. 1987; Swanson et al. 1987) and in situ hybridiza-
ules, which they suggest better reflect “drug craving” or tion (Dineley-Miller and Patrick 1992; Flores et al. 1992;
incentive motivation (Markou et al. 1993). Their finding Wada et al. 1989). Due to the wide distribution of nAChRs
was that while that there was no significant sex differ- nicotine addiction or chronic administration may alter
ence in the fixed-ratio schedule, females had higher the function of most, if not all, of the brain regions and
breaking points in the progressive-ratio schedule. The various neurotransmitter systems.
females, in both contingency paradigms, also demon- Although nAChRs are located in many peripheral tis-
strated a shorter latency to first infusion, and this may be sues, the rewarding properties of nicotine are mediated
interpreted as another index reflecting nicotine incentive by the interaction of this alkaloid with nAChRs in the
motivation. CNS. In animals, systemic administration of a nicotinic
Lastly, in an attempt to look at the interplay between antagonist which crosses the blood-brain barrier (meca-
genetics and the SA of drugs of abuse, Todte et al. (2001) mylamine) attenuates the intravenous SA of nicotine
examined the propensity of Brown-Norway rats (BNR) (Corrigall and Coen 1989; Goldberg et al. 1981; Risner
and Wistar-Kyoto rats (WKR) to orally self-administer and Goldberg 1983). Similarly, mecamylamine pretreat-
ethanol, nicotine, cocaine, and morphine. The WKR ment blocks the effects of nicotine in a conditioned place
strain is known for its higher emotionality and its lower preference assay (Fudala et al. 1985) and inhibits the
plasma stress catecholamine responses (Todte et al. stimulant effect of nicotine on locomotor activity, in
2001). They also examined age as an important mediating chronically treated rats (Benwell et al. 1995). During
108

precipitated nicotine withdrawal, recent data demonstrate when nicotine is substituted for cocaine (Epping-Jordan
that central nAChRs mediate elevated brain stimulation et al. 1999; Picciotto et al. 1998). Similar studies with
reward thresholds, whereas both central and peripheral α4- or α7-deficient and viable mice have not been de-
nAChRs appear to mediate the somatic withdrawal syn- scribed to date. Nicotine-induced place preference exper-
drome (Watkins et al. 2000a, 2000b). iments using these three types of genetically altered mice
Neuronal nAChRs are pentameric and allosteric have also not been documented, although a recent report
membrane protein complexes. They are members of the suggests that the β2 subunit contributes to the motiva-
superfamily of ligand-gated ion channels permeable to tional effect of cocaine (Zachariou et al. 2001). Taken
cations (Na+/K+, Ca2+; see Changeux et al. 1984), and together, it is therefore possible that the reinforcing and
this ionic flux leads to a depolarization and an excitatory motivational properties of nicotine are mediated by a
response of neurons. Each neuronal nAChR subunit is neuronal nAChR made of at least one β2 subunit.
encoded by a separate gene. To date, of the 12 distinct In contrast to many other receptors that are down-
genes cloned, the expression of nine has been found in regulated following a chronic exposure to their agonist,
the brain of rat or primate. These subunits have been nAChR are unexpectedly up-regulated after chronic
named α2, α3, α4, α5, α6, α7, β2, β3, and β4. There- nicotine administration (Marks et al. 1983; Schwartz and
fore, multiple combinations of some but not all of these Kellar 1983; for a review see Wonnacott 1990). Instead
five subunits can form functional heteromeric (α with β) of a decrease in the number of nicotine binding sites, an
or homomeric (α) nAChRs with distinct pharmacophysi- increase following chronic nicotine administration has
ological features (Lukas et al. 1999; Zoli et al. 1998). In been observed in postmortem brain studies of rodents
addition to this molecular complexity, the regional distri- (Marks et al. 1992; Pauly et al. 1991, 1996; Zhang et al.
bution of the different subunits differs among brain areas 1994) and humans (Benwell et al. 1988; Breese et al.
and among neuronal populations. In brain, the most 1997; Court et al. 1998) as well as in the living brain of
widely and concurrently expressed subunits are α4 and baboons (Kassiou et al. 2001). Chronic exposure of rats
β2. In contrast, the patterns of α7, α3, and α5 subunit to mainstream cigarette smoke for more than 3 months
expression are more limited in both their density and resulted in an increase in nicotine binding sites in the
localization. The α2, α6, β3, and β4 subunits are limited cerebral and cerebellar cortices and more modestly in the
to only a few brain regions. Although the molecular striatum but not in the thalamus or hippocampus where
diversity of the initial site of action for nicotine is well nAChR are also densely expressed (Yates et al. 1995).
explored, it is still unknown whether one (or several) With few exceptions studies using multiple injections of
particular subunit combination of nAChR preferentially various doses of nicotine have shown increases in nico-
mediates the reinforcing effects of nicotine, or whether a tine binding sites in the cortex, the striatum, and the
particular genotype(s) is involved in tobacco consump- hippocampus (Collins et al. 1988; Coutcher et al. 1992;
tion behavior. Studies with subtype-selective nAChR el-Bizri and Clarke 1994; Flores et al. 1997; Ksir et al.
antagonists, focusing on their ability to modulate nicotine’s 1987; Lapchak et al. 1989; Nordberg et al. 1989; Romanelli
reinforcing (and other) effects would therefore be of et al. 1988; Rowell and Li 1997; Ulrich et al. 1997; Wall
value. For example, a recent study has shown that the et al. 2000; Zhang et al. 1994). The up-regulation of
α4/β2 nAChR selective antagonist erysodine blocked the nicotine binding sites, which was also found in the
discriminative and reinforcing effects of nicotine in rats brains of rats that self-administered nicotine (Donny et
(Mansbach et al. 2000). al. 2000), occurs largely in specific cerebral regions.
Studies using transgenic mice that lack specific sub- This may reflect a differential responsiveness of nAChR
units of the nAChR could also help determine which to chronic exposure to nicotine, possibly related to a dif-
subunits play a fundamental role in nicotine addiction ferential subunit composition. Immunoprecipitation pro-
(for reviews see Cordero-Erausquin et al. 2000; Marubio tocols or binding studies using specific subtype ligands
and Changeux 2000; Picciotto et al. 2000). Several (but suggest that the subunit composition of the up-regulated
not all possible) mice lacking one or two subunits have nAChRs is made up of α4 and β2 subunits in the cortex
been generated. While some of these appear to have and the hippocampus of rats chronically treated with
multiorgan dysfunctions and impaired survival, others nicotine and not of α3 or β4 subunits (Flores et al. 1992,
reach adulthood (Bansal et al. 2000; Caldarone et al. 1997). Chronic nicotine exposure in mice also led to an
2000; Franceschini et al. 2000; Marubio et al. 1999; Paylor increase in α-bungarotoxin sites (thought to represent
et al. 1998; Ross et al. 2000; Rossi et al. 2001; Vetter et nAChR that contain the α7 subunit; Marks et al. 1983;
al. 1999; Xu et al. 1999a, 1999b). Their behavior has not Pauly et al. 1991), although this has not been found in
been fully explored in the context of nicotine reinforce- the postmortem brains of human smokers (Court et al.
ment. While α7 subunit-deficient mice have no detect- 1998).
able abnormalities in the density of [3H]nicotine sites It has been proposed that simultaneously with an
(Orr-Urtreger et al. 1997), α4 and β2-subunit deficient increased number of nicotine sites a functional desensiti-
mice have lost almost all their sites in all brain regions zation of the nAChR occurs during chronic exposure to
(Marubio et al. 1999; Ross et al. 2000). Interestingly, β2 nicotine. Indeed, the nAChR is an allosteric protein com-
subunit-deficient mice which have been trained to self- plex with at least three conformational states. The inter-
administer cocaine cease their nose-poking behavior action of nicotine or acetylcholine (low affinity) with the
109

nAChR induces a quick transition from the resting state Downstream sites of action: the mesolimbic
(channel closed) to the active state (channel opened with dopamine neurons
high affinity). If nicotine exposure is prolonged, the
receptor slowly enters into a desensitized state (channel Although the function of numerous neurotransmitter
closed) that is unresponsive to the agonist. The in- systems expressing nAChR could be altered by nicotine,
creased number of nicotine sites could be a neuronal the initial reinforcing properties of this compound are
response to the accumulation of receptors desensitized thought to be mediated primarily through the mesolimbic
after long-term nicotine exposure. Desensitization of and mesocortical dopamine (DA) systems. These neuro-
nAChR populations may be relevant in human smokers, nal pathways consist primarily of the DA neurons from
since prolonged exposure to low nicotine concentrations the ventral tegmental area (VTA) that project to various
(such as may be observed in “trough” periods or at cerebral regions, notably the nucleus accumbens and
nighttime) may result in receptor desensitization (see various limbic cortices. These brain areas are involved in
“Chronopharmacokinetic characteristics of nicotine” reward functions and mediate the actions of natural
above; see also Marks et al. 1996; Rosecrans and Karan reinforcers as well as many drugs of abuse.
1993) To our knowledge, SA of nicotine directly into those
In mice regardless of the nicotine administration pro- DA terminal areas has not been described to date. How-
tocol used changes in nAChR numbers are not attribut- ever, direct administration of nicotine in the mesolimbic
able to transient or sustained changes in nAChR subunit and mesocortical regions substitutes for the discrimina-
mRNA levels (Pauly et al. 1996). The nature of the tive stimulus of nicotine established through subcutane-
mechanisms involved in the up-regulation of nicotine ous injection (Miyata et al. 1999). The importance of
receptors is therefore likely posttranscriptional, although those regions, especially the nucleus accumbens, was
increased synthesis may not be totally excluded (Ryan also supported by the finding that bilateral lesions of
and Loiacono 2001; Yu et al. 1996). DA neurons led to a reduction in intravenous nicotine
Since the reported up-regulation of nAChR is also SA (Corrigall et al. 1992; Singer et al. 1982). A bilateral
accompanied by enhanced behavioral responses such as intra-accumbens microinjection of 6-hydroxydopamine
locomotor activity or cognitive performance, this molec- into the accumbens also abolished the locomotor stimu-
ular event might be related to the behavioral sensitiza- lant effect of subcutaneous nicotine in rats (Clarke et al.
tion observed during chronic nicotine exposure (Abdulla 1988). Similarly, the blockade of DA transmission by
et al. 1996; Ksir et al. 1985, 1987; Nisell et al. 1996). Of selective DA antagonists in nonlesioned rats also
greater salience to the field of nicotine addiction is the reduced nicotine SA (Corrigall and Coen 1991b) and
question of whether changes in the density of nicotinic completely blocked nicotine-conditioned place prefer-
sites are a crucial molecular event in the early abstinence ence (Acquas et al. 1989) and nicotine-induced ICSS
phase. It could be hypothesized that as soon as the con- (Huston-Lyons et al. 1993). These observations, in sum,
centration of nicotine is no longer sufficient to desensi- suggest a major role for the DA neurons, especially
tize nAChR, the desensitized nAChR return to their rest- from the mesolimbic system, in mediating the reinforcing
ing state. Consequently, when nicotine exposure is termi- effects of nicotine.
nated, the total number of functional nAChR potentially Nicotinic AChRs are localized on the cell bodies of
sensitive to nicotine (or acetylcholine) should be greatly DA neurons in the ventral mesencephalon as well as on
increased at the cell surface, thus rapidly altering the DA terminals in the nigrostriatal and mesolimbic path-
physiology of neuronal circuits. However, in adult ways (Clarke and Pert 1985; Clarke et al. 1984; London
rodents the return of the up-regulated receptors to basal et al. 1985). Most of those DA neurons express abun-
values takes at least 1 week after the cessation of nico- dantly the α6 and β3 nAChR subunit mRNA (Le Novere
tine administration (although not in all brain regions; et al. 1996) and also at moderate levels α3, α4, α5, and
Collins et al. 1990; Koylu et al. 1997; Pietila et al. 1998). β2 subunit mRNA (Dineley-Miller and Patrick 1992;
In addition, a more robust and longer lasting receptor Duvoisin et al. 1989; Sorenson et al. 1998; Wada et al.
up-regulation was observed in rats exposed to nicotine 1989). Although there is no unique expression pattern of
during the adolescent period (Trauth et al. 1999). Overall, these various nAChR subunits within DA neurons from
it is possible that the persistent up-regulation of nicotine the VTA, the activation of nAChR from this area alone
binding sites during cessation may play a role in the could explain some of the behavioral effects seen after
intensity of the early withdrawal symptoms and therefore repeated systemic administration of nicotine. Indeed,
the likelihood of relapse. infusion of a nicotinic antagonist into the VTA prior to
Knowledge of the neurobiological and behavioral intravenous nicotine SA sessions significantly decreases
consequences of the increase in nicotine binding sites in the number of nicotine infusions obtained (Corrigall et
various brain regions is therefore important to extend our al. 1994). Moreover, repeated applications of a nicotinic
understanding of altered function in active smokers as agonist (cytisine) directly into the VTA induced a place
well as in others exposed to nicotine (such as nonsmokers, preference for the cytisine paired compartment (Museo
newborns exposed during pregnancy, and ex-smokers and Wise 1994), and repeated applications of nicotine
therapeutically exposed to nicotine through nasal spray, showed a progressive increase in the locomotor response
transdermal patches, or gum). in rats (Panagis et al. 1996; Reavill and Stolerman 1990).
110

These data support the hypothesis that the α3, α4, α5, Nicotinic receptors localized on DA cell bodies or
α6, β2, and β3 subunit containing nAChR, which are terminals could initiate nicotine-induced DA release. In
localized on the mesolimbic system, contribute to both fact, in vivo microdialysis studies have shown that the
the reinforcing properties and the psychostimulant local infusion of nicotine directly into the VTA is suffi-
effects of nicotine. cient to increase the DA extracellular concentration in
The interaction of nicotine with nAChR localized on the nucleus accumbens (Marshall et al. 1997; Mifsud et
the VTA-accumbens pathway causes profound modifica- al. 1989; Nisell et al. 1994). However, although persis-
tions in the functioning of DA neurons. Acute as well as tent effects are observed when nicotine is administered
chronic intermittent administration of nicotine leads to locally to the DA cell bodies, local administration to the
an increase in the activity of midbrain DA neurons by terminal fields of DA neurons can also stimulate (albeit
accelerating the firing of neurons and by switching their transiently) DA release (Mifsud et al. 1989; Nisell et al.
activity from regular to burst firing patterns (Grenhoff et 1994). VTA or somatodendritic nAChRs have been
al. 1986; Lichtensteiger et al. 1982; Mereu et al. 1987; implicated in the effects of nicotine in SA, conditioned
Nisell et al. 1996). Moreover, cessation of continuous place preference assays, or on spontaneous activity. Thus
nicotine administration by subcutaneous osmotic mini- nAChRs of the VTA have a crucial role in both nicotine-
pump (6 mg/kg for 12 days) also alters firing rates of induced DA release and in the reinforcing properties of
midbrain DA neurons (Rasmussen and Czachura 1995). nicotine.
These changes in electrophysiological activity, especially A possibly contrasting line of evidence involves the
burst firing, could be associated with greater neurotrans- finding that schizophrenic patients receiving haloperidol
mitter release in the terminal fields of DA neurons, initi- (a dopaminergic D2-like antagonist) smoke at high rates
ating neurochemical changes within those neurons or (e.g., McEvoy 1995). It is possible that this behavior
within neurons from projection areas. continues because high levels of tobacco use may allow
An acute systemic injection of nicotine, such as with the individual to compensate for decreases in dopamine-
other drugs of abuse, increases the extracellular concen- based reinforcement caused by haloperidol (see also
trations of DA in the terminal fields of the mesolimbic Rose and Corrigal 1997) and because the individual may
system (Imperato et al. 1986; Maisonneuve and Glick be receiving continued reinforcement from nonnicotine
1999; Mifsud et al. 1989; Pontieri et al. 1996; for a stimuli involved in smoking.
review see Di Chiara 2000), the nigrostriatal pathway Several investigators have also suggested that the
(Di Chiara and Imperato 1988; Toth et al. 1992), and the maintenance of drug SA (and particularly nicotine or
mesocortical pathway (Bassareo et al. 1996; Summers tobacco) may also depend critically on cues associated
and Giacobini 1995; Toth et al. 1992). Repeated nicotine with smoking, which may acquire “conditioned reinforcer”
injections in rats (0.4 mg/kg nicotine tartrate subcutane- status by their pairing with self-administered nicotine
ously for 7 days) increased the dopamine release caused and its secondary dopamine release (Balfour et al. 2000;
by local nicotine administration in the striatum, as mea- Di Chiara 2000; Donny et al. 1999; see also above). Of
sured with in vivo microdialysis (Marshall et al. 1997). particular relevance to this hypothesis is the temporally
Interestingly, this effect was not observed if the same contingent relationship between smoking behavior (e.g.,
daily nicotine dose was administered as a continuous in- puff inhalation) and its associated sensory stimuli or
fusion (Marshall et al. 1997). In other experiments with “cues”, with a rapid rise in intra-arterial nicotine levels
chronically treated rats, a systemic injection of nicotine (see “Chronopharmacokinetic characteristics of nicotine”,
increased DA release in both the prefrontal cortex (Nisell above). It has also been suggested that the maintenance
et al. 1996) and the nucleus accumbens (Benwell et al. of nicotine or tobacco SA is especially dependent on
1994; Nisell et al. 1997; Reid et al. 1996). The nicotine- possible secondary reinforcers, due to the presumed
induced DA release is of the same amplitude or higher in desensitization of nAChRs after prolonged exposure
chronically (as compared to acutely) injected rats. Rats (Balfour et al. 2000).
treated daily with nicotine may have both higher basal
extracellular DA concentrations in the nucleus accumb-
ens, as well as an increased DA release in response to a The endogenous opioid system and its connection
further nicotine challenge (Benwell and Balfour 1992; to the withdrawal process
Benwell et al. 1994, 1995; Damsma et al. 1989). Local
intrategmental mecamylamine injection reduced DA Preclinical studies on nicotine reinforcement exploring
overflow in the ipsilateral nucleus accumbens of rats the involvement of the endogenous opioid system –
treated chronically with nicotine (by osmotic minipump) µ-, δ-, or κ-opioid receptors or their endogenous agonist
for 14 days (Hildebrand et al. 1999). A similar effect was peptides, the enkephalins, dynorphins, and β-endorphin
observed after the termination of nicotine administration – appear to be rare. In addition, in most animal studies
(Fung et al. 1996). These observations suggest that no conclusive involvement has been found (for a review
dynamic DA transmission is preserved (and may be see Pomerleau 1998). For example, a preinjection with
enhanced) in the mesolimbic and corticolimbic systems κ/µ-opioid receptor antagonist naltrexone prior to SA
during intermittent nicotine administration, and may be sessions did not affect the reinforcing properties of nicotine
decreased during withdrawal. (Corrigall and Coen 1991b), and an injection of naloxone
Table 1 Early to late effects subsequent to a chronic (intermittent or continuous) nicotine administration on the endogenous opioid system in the male rodent brain

POMC mRNA End peptide ppEnk mRNA Met-Enk peptide ppDyn mRNA [3H]-DAMGO
site Bmax and Kd

POMC system
Chronic nicotine effect (<2 h) =Hypothalamus,
anterior pituitaryb;
↑anterior pituitaryb;
↓intermediate pituitaryb
Early withdrawal (4 h–≤3 d) ↓Hypothalamusf ↓Hypothalamusg;
=pituitaryg
Late withdrawal (7 d–21 d) =Hypothalamusf; ↑Hypothalamusg
↓hypothalamusg;
↓hypothalamusg
Enkephalin system
Chronic nicotine effect (≤2 h) =Striatum, accumbens, =Hippocampus, cortexa;
hippocampus, hypothalamus, ↑Hypothalamusa;
cerebellumb,c,e; ↓striatum, midbrainh
↓striatum, hippocampusc
Early withdrawal (24 h–≤3 d) =Striatum, accumbens, =Striatum, hippocampus,
cerebellumc,d; ↑striatum, hypothalamus, pituitaryg;
accumbens, hippocampusc,d,f; ↑striatuma
↓striatumc
Late withdrawal (7 d–14 d) =Striatum, hippocampus, =Striatum, hippocampus,
cerebellumc hypothalamus, pituitaryf
Dynorphin system
Chronic nicotine effect (≤2 h) =Striatum, accumbens,
hippocampus,
hypothalamusb,e;
↓ventral accumbense
Early withdrawal (24 h) =Striatum, accumbensd
Mu receptor
Chronic nicotine effect (≤2 h) =Striatum,
midbraing
a Dhatt et al. (1995) d Mathieu et al. (1996) g Rosecrans et al. (1985)
b Hollt and Horn (1992) e Mathieu-Kia and Besson (1998) h Wewers et al. (1999)
c Houdi et al. (1998) f Rasmussen (1998)
111
112

did not dose-dependently block the low reward threshold rodents was also affected during the early and late phases
induced by nicotine in the ICSS assay (Huston-Lyons of nicotine withdrawal (Table 1). For example, 4 h after
and Kornetsky 1992). Tests of generalization to mor- the end of a 4-week period of nicotine administration,
phine in drug-discrimination assays were also negative the expression of the pro-opiomelanocortin (POMC)
(Stolerman 1989). Other data show that the selective gene, the biosynthetic precursor for β-endorphin and
δ-opioid receptor antagonist naltrindole, given in combi- ACTH, was decreased in the hypothalamus (Rasmussen
nation with nicotine for 3 days, failed to prevent the 1998). This effect persisted for at least 3 days following
expression and the development of sensitization to the the end of nicotine administration and after a transient
locomotor-activating effects of nicotine (Heidbreder et normalization, decreased again 3 weeks later (Rasmussen
al. 1996). While opioid antagonists have been shown to 1998). Therefore chronic intermittent nicotine exposure
have no direct impact on the effects of nicotine, studies induced a biphasic, long-lasting inhibition of POMC
with opioid agonists suggest a possible subtype-selective gene expression in the hypothalamus. It is therefore pos-
role. For example, the κ-opioid agonist U69593 did not sible that there are similar decreases in the biosynthesis
prevent the sensitized motor responses observed after of POMC-derived peptides as well (e.g., β-endorphin).
repeated nicotine administration (Heidbreder et al. 1995). For example, in mice, hypothalamic immunoreactive
However, a selective µ-opioid agonist [(D-Ala2, MePhe4- β-endorphin is reduced 24 h and 7 days following a
Gly5-ol) enkephalin] administered into the VTA produced chronic intermittent nicotine administration but is
a reduction in nicotine SA in a reinstatement model increased at 2 weeks (Rosecrans et al. 1985). In contrast,
(Corrigall et al. 2000). in the pituitary gland, levels of β-endorphin immunore-
In humans, it has been reported that the administra- activity remain unaltered during nicotine withdrawal
tion of the opioid antagonist naloxone could reduce the (Rosecrans et al. 1985). No information is currently
number of cigarettes smoked by chronic smokers (Gorelick available on the pituitary level of POMC during early
et al. 1988) and block the subjective pleasurable response and late nicotine withdrawal.
to smoking, thus suggesting that opioid peptides are The expression of the enkephalins may also be modu-
released during smoking and may be partially responsible lated by chronic nicotine exposure. Several rodent studies
for smoking maintenance. In turn, most of the withdrawal have shown altered levels of these opioid peptides (Dhatt
symptoms are not affected by naltrexone in heavy smok- et al. 1995; Pierzchala et al. 1987; Van Loon et al. 1991;
ers, although naltrexone appears to reduce the perceived Wewers et al. 1999), but changes were interpreted mostly
difficulty of abstaining during the first 24 h of cigarette as a short-term effect of the chronic nicotine regimen
abstinence (Sutherland et al. 1995). rather than as a consequence of withdrawal. However,
In animal models, naloxone precipitates not only an some studies have reported that administration of nico-
ICSS threshold elevation in rats that received chronic tine for 2 weeks produced an increase in preproenkephalin
continuous nicotine but also a conditioned place aversion mRNA levels 24 h following the final injection. Prepro-
(Watkins et al. 2000a, 2000b). Also, as noted above, the enkephalin is the peptide precursor for Leu- and Met-
physical signs induced by cessation of nicotine exposure enkephalins, two biologically active opioid peptides
resemble those of the rat opiate withdrawal syndrome found in both the striatum (Dhatt et al. 1995; Houdi et al.
(Malin et al. 1993; Watkins et al. 2000b). After chronic 1998) and the nucleus accumbens (e.g., Mathieu et al.
nicotine exposure, a physical withdrawal syndrome 1996). Similar increases have also been described in the
could be precipitated in rats using injections of either hippocampus (Houdi et al. 1998). In contrast, no change
nAChR or opioid receptor antagonists (Hildebrand et al. in enkephalin levels was observed in those brain regions
1997, 1999; Malin et al. 1993, 1994; Watkins et al. within the 2 h following the last nicotine injection of a
2000b). Also, naloxone initiates a withdrawal syndrome chronic regimen (Hollt and Horn 1992; Houdi et al.
in rats that was reversible by a subsequent morphine 1998; Mathieu-Kia and Besson 1998; Mathieu-Kia et al.
administration (Malin et al. 1992). Similarly, the systemic 2000). This suggests that enkephalins are involved
administration of an analog of neuropeptide FF, which primarily in the withdrawal syndrome (which may take
precipitates withdrawal symptoms in morphine-dependent 24 h to appear following the end of nicotine administra-
rats, also precipitated a withdrawal syndrome in nicotine- tion) rather than in the processes involved in ongoing
dependent rats (Malin et al. 1996). Interestingly, meca- nicotine exposure.
mylamine-precipitated nicotine withdrawal was attenuated Less information is available on the influence of nico-
by preinjection with the µ-opioid receptor agonist tine exposure on the dynorphins, although they are a
morphine as well as with the selective δ-opioid agonist major neurochemical substrate in the DA terminal fields,
TAN-67 (Ise et al. 2000). One possible hypothesis is that and are altered by the effects of psychostimulant drugs
chronic intermittent nicotine exposure results in a such as cocaine or amphetamine. Available studies
prolonged release of endogenous opioid peptides, and indicate that dynorphin synthesis is unaltered in many
that nicotine withdrawal therefore decreases tone in regions of the rodent brain after repeated nicotine expo-
opioid receptor systems. sure (Hollt and Horn 1992; Mathieu-Kia and Besson
In addition to the behavioral evidence of an interac- 1998; Mathieu et al. 1996; Mathieu-Kia et al. 2000), so it
tion between nicotinic and opioid systems during with- is therefore premature to postulate a role for dynorphin
drawal, the hypothalamic β-endorphinergic system in synthesis in the different phases of nicotine withdrawal.
113

Overall, a clear relationship between the endogenous Balfour DJK, Wright AE, Benwell MEM, Birrell CS (2000) The
opioids and behaviors observed during nicotine cessation putative role of extra-synaptic mesolimbic dopamine in the
neurobiology of nicotine dependence. Behav Brain Res 113:
is not currently evident. Early signs such as the persis- 73–83
tent drug-seeking behavior observed during saline substi- Bansal A, Singer JH, Hwang BJ, Xu W, Beaudet A, Feller MB
tution, the somatic withdrawal signs, the changes in (2000) Mice lacking specific acetylcholine receptor subunits
ICSS thresholds or measure of anxiogenic responses exhibit dramatically altered spontaneous activity patterns and
reveal a limited role for retinal waves in forming ON and OFF
culminate within a 4-day period after nicotine exposure. circuits in the inner retina. J Neurosci 20:7672–7681
These behavioral observations may be associated with Bardo MT, Green TA, Crooks PA, Dwoskin LP (1999) Nornicotine
changes in the β-endorphinergic or enkephalinergic is self-administered intravenously by rats. Psychopharmacology
systems. 146:290–296
Bassareo V, Tanda G, Petromilli P, Giua C, Di Chiara G (1996)
Non-psychostimulant drugs of abuse and anxiogenic drugs
activate with differential selectivity dopamine transmission in
Conclusions and perspectives the nucleus accumbens and in the medial prefrontal cortex of
the rat. Psychopharmacology 124:293–299
Animal studies have addressed questions related to the Bauco P, Wise RA (1994) Potentiation of lateral hypothalamic and
midline mesencephalic brain stimulation reinforcement by
reinforcing effects of nicotine, and have established nicotine: examination of repeated treatment. J Pharmacol Exp
some of the molecular and neuronal mechanisms under- Ther 271:294–301
lying chronic nicotine use. However, our experimental Benowitz NL (1988) Drug therapy. Pharmacologic aspects of
approaches for understanding nicotine addiction may not cigarette smoking and nicotine addition. N Engl J Med 319:
1318–1330
globally model tobacco smoking behavior observed in Benowitz NL, Kuyt F, Jacob P, III (1982) Circadian blood nicotine
humans. Animal treatment protocols with long and fre- concentrations during cigarette smoking. Clin Pharmacol Ther
quent access to the drug may reproduce the circadian 32:758–764
oscillation of nicotine concentrations in the body and the Benowitz NL, Perez-Stable EJ, Fong I, Modin G, Herrera B,
brain of smokers. The repetitive “spiking” and the gradual Jacob P, III (1999) Ethnic differences in N-glucuronidation of
nicotine and cotinine. J Pharmacol Exp Ther 291:196–1203
“ceiling” profile of drug use throughout the day also Benwell ME, Balfour DJ (1992) The effects of acute and repeated
needs to be incorporated in future research. In addition, nicotine treatment on nucleus accumbens dopamine and
we may be entering into a new era in which smokers use locomotor activity. Br J Pharmacol 105:849–856
cigarettes in a more constrained pattern as a consequence Benwell ME, Balfour DJ, Anderson JM (1988) Evidence that
tobacco smoking increases the density of (–)-[3H]nicotine
of legal restrictions and social pressure. These new binding sites in human brain. J Neurochem 50:1243–1247
factors must be considered in the development of new Benwell ME, Balfour DJ, Khadra LF (1994) Studies on the
animal models of nicotine addiction. Experimental strat- influence of nicotine infusions on mesolimbic dopamine and
egies to model withdrawal/abstinence and relapse could locomotor responses to nicotine. Clin Investig 72:233–239
Benwell ME, Balfour DJ, Birrell CE (1995) Desensitization of the
also be used to investigate the neuronal substrates that nicotine-induced mesolimbic dopamine responses during
could serve as targets for the development of novel constant infusion with nicotine. Br J Pharmacol 114:454–460
pharmacotherapies for nicotine addiction. Bespalov A, Lebedev A, Panchenko G, Zvartau E (1999) Effects
of abused drugs on thresholds and breaking points of intra-
Acknowledgements This work was supported in part by the NIH- cranial self-stimulation in rats. Eur Neuropsychopharmacol 9:
NIDA Grants DA05130, DA00049, and DA11113. The authors are 377–383
grateful to Dr. J. Mantsch for valuable and helpful discussions el-Bizri H, Clarke PB (1994) Regulation of nicotinic receptors in
concerning many aspects of this manuscript. rat brain following quasi-irreversible nicotinic blockade by
chlorisondamine and chronic treatment with nicotine. Br
J Pharmacol 113:917–925
Bozarth MA, Pudiak CM, KuoLee R (1998a) Effect of chronic
References nicotine on brain stimulation reward. I. Effect of daily injections.
Behav Brain Res 96:185–188
Abdulla FA, Bradbury E, Calaminici MR, Lippiello PM, Wonnacott Bozarth MA, Pudiak CM, KuoLee R (1998b) Effect of chronic
S, Gray JA, Sinden JD (1996) Relationship between up- nicotine on brain stimulation reward. II. An escalating dose
regulation of nicotine binding sites in rat brain and delayed regimen. Behav Brain Res 96:189–194
cognitive enhancement observed after chronic or acute nicotinic Breese CR, Marks MJ, Logel J, Adams CE, Sullivan B,
receptor stimulation. Psychopharmacology 124:323–331 Collins AC, Leonard S (1997) Effect of smoking history on
Acquas E, Carboni E, Leone P, Di Chiara G (1989) SCH 23390 [3H]nicotine binding in human postmortem brain. J Pharmacol
blocks drug-conditioned place-preference and place-aversion: Exp Ther 282:7–13
anhedonia (lack of reward) or apathy (lack of motivation) after Buczek Y, Lê AD, Stewart J, Shaham Y (1999) Stress reinstates
dopamine-receptor blockade? Psychopharmacology 99:151–155 nicotine seeking but not sucrose solution seeking in rats.
Acri JB, Grunberg NE, Morse DE (1991) Effects of nicotine on Psychopharmacology 144:183–188
the acoustic startle reflex amplitude in rats. Psychopharmacology Calcagnetti DJ, Schechter MD (1994) Nicotine place preference
104:244–248 using the biased method of conditioning. Prog Neuropsycho-
Adir J, Miller RP, Rotenberg KS (1976) Disposition of nicotine in pharmacol Biol Psychiatry 18:925–933
the rat after intravenous administration. Res Commun Chem Caldarone BJ, Duman CH, Picciotto MR (2000) Fear conditioning
Pathol Pharmacol 13:173–183 and latent inhibition in mice lacking the high affinity subclass
Ando K, Yanagita T (1981) Cigarette smoking in rhesus monkeys. of nicotinic acetylcholine receptors in the brain. Neuropharma-
Psychopharmacology 72:117–127 cology 39:2779–2784
Balfour DJ (1991) The influence of stress on psychopharmacological Carboni E, Bortone L, Giua C, Di Chiara G (2000) Dissociation
responses to nicotine. Br J Addict 86:489–493 of physical abstinence signs from changes in extracellular
114
dopamine in the nucleus accumbens and in the prefrontal cortex acetylcholine receptor in mammalian brain. Proc Natl Acad
of nicotine dependent rats. Drug Alcohol Depend 58:93–102 Sci USA 84:8697–8701
Carroll ME, Lac ST, Asencio M, Keenan RM (1989) Nicotine Dhatt RK, Gudehithlu KP, Wemlinger TA, Tejwani GA, Neff NH,
dependence in rats. Life Sci 45:1381–1388 Hadjiconstantinou M (1995) Preproenkephalin mRNA and
Changeux JP, Devillers-Thiery A, Chemouilli P (1984) Acetylcho- methionine-enkephalin content are increased in mouse striatum
line receptor: an allosteric protein. Science 225:1335–1345 after treatment with nicotine. J Neurochem 64:1878–1883
Chiamulera C, Borgo C, Falchetto S, Valerio E, Tessari M (1996) Di Chiara G (2000) Role of dopamine in the behavioural actions
Nicotine reinstatement of nicotine self-administration after of nicotine related to addiction. Eur J Pharmacol 393:295–314
long-term extinction. Psychopharmacology 127:102–107 Di Chiara G, Imperato A (1988) Drugs abused by humans prefer-
Clarke PBS, Fibiger HC (1987) Apparent absence of nicotine- entially increase synaptic dopamine concentrations in the
induced conditioned place preference in rats. Psychopharma- mesolimbic system of freely moving rats. Proc Natl Acad Sci
cology 92:84–88 USA 85:5274–5278
Clarke PB, Pert A (1985) Autoradiographic evidence for nicotine Dineley-Miller K, Patrick J (1992) Gene transcripts for the nico-
receptors on nigrostriatal and mesolimbic dopaminergic neurons. tinic acetylcholine receptor subunit, beta4, are distributed in
Brain Res 348:355–358 multiple areas of the rat central nervous system. Brain Res
Clarke PB, Pert CB, Pert A (1984) Autoradiographic distribution Mol Brain Res 16:339–344
of nicotine receptors in rat brain. Brain Res 323:390–395 Djurič VJ, Dunn E, Overstreet DH, Dragomir A, Steiner M (1999)
Clarke PB, Schwartz RD, Paul SM, Pert CB, Pert A (1985) Nico- Antidepressant effect of ingested nicotine in female rats of
tinic binding in rat brain: autoradiographic comparison of Flinders resistant and sensitive lines. Physiol Behav 67:533–537
[3H]acetylcholine, [3H]nicotine, and [125I]-alpha-bungarotoxin. Donny EC, Caggiula AR, Knopf S, Brown C (1995) Nicotine self-
J Neurosci 5:1307–1315 administration in rats. Psychopharmacology 122:390–394
Clarke PB, Fu DS, Jakubovic A, Fibiger HC (1988) Evidence that Donny EC, Caggiula AR, Mielke MM, Jacobs KS, Rose C,
mesolimbic dopaminergic activation underlies the locomotor Sved AF (1998) Acquisition of nicotine self-administration in
stimulant action of nicotine in rats. J Pharmacol Exp Ther 246: rats: the effects of dose, feeding schedule, and drug contingency.
701–708 Psychopharmacology 136:83–90
Collins AC, Romm E, Wehner JM (1988) Nicotine tolerance: Donny EC, Caggiula AR, Mielke MM, Booth S, Gharib MA,
an analysis of the time course of its development and loss in Hoffman A, Maldovan V, Shupenko C, McCallum SE (1999)
the rat. Psychopharmacology 96:7–14 Nicotine self-administration in rats on a progressive ratio
Collins AC, Bhat RV, Pauly JR, Marks MJ (1990) Modulation of schedule of reinforcement. Psychopharmacology 147:135–142
nicotine receptors by chronic exposure to nicotinic agonists Donny EC, Caggiula AR, Rowell PP, Gharib MA, Maldovan V,
and antagonists. Ciba Found Symp 152:68–82 Booth S, Mielke MM, Hoffman A, McCallum S (2000)
Cordero-Erausquin M, Marubio LM, Klink R, Changeux JP Nicotine self-administration in rats: estrous cycle effects, sex
(2000) Nicotinic receptor function: new perspectives from differences and nicotinic receptor binding. Psychopharmacology
knockout mice. Trends Pharmacol Sci 21:211–217 151:392–405
Corrigall WA, Coen KM (1989) Nicotine maintains robust self- Duvoisin RM, Deneris ES, Patrick J, Heinemann S (1989) The func-
administration in rats on a limited-access schedule. Psycho- tional diversity of the neuronal nicotinic acetylcholine receptors
pharmacology 99:473–478 is increased by a novel subunit: beta 4. Neuron 3:487–496
Corrigall WA, Coen KM (1991a) Opiate antagonists reduce cocaine Emmett-Oglesby MW, Mathis DA, Moon RT, Lal H (1990)
but not nicotine self-administration. Psychopharmacology 104: Animal models of drug withdrawal symptoms. Psychopharma-
167–170 cology 101:292–309
Corrigall WA, Coen KM (1991b) Selective dopamine antagonists Epping-Jordan MP, Watkins SS, Koob GF, Markou A (1998)
reduce nicotine self-administration. Psychopharmacology 104: Dramatic decreases in brain reward function during nicotine
171–176 withdrawal. Nature 393:76–79
Corrigall WA, Franklin KB, Coen KM, Clarke PB (1992) The Epping-Jordan MP, Picciotto MR, Changeux JP, Pich EM (1999)
mesolimbic dopaminergic system is implicated in the reinforcing Assessment of nicotinic acetylcholine receptor subunit contri-
effects of nicotine. Psychopharmacology 107:285–289 butions to nicotine self-administration in mutant mice. Psycho-
Corrigall WA, Coen KM, Adamson KL (1994) Self-administered pharmacology 147:25–26
nicotine activates the mesolimbic dopamine system through Faraday MM, Scheufele PM, Rahman MA, Grunberg NE (1999)
the ventral tegmental area. Brain Res 653:278–284 Effects of chronic nicotine administration on locomotion
Corrigall WA, Coen KM, Adamson KL, Chow BL, Zhang J depend on rat sex and housing condition. Nicotine Tob Res 1:
(2000) Response of nicotine self-administration in the rat to 143–155
manipulations of mu-opioid and gamma-aminobutyric acid Flores CM, Rogers SW, Pabreza LA, Wolfe BB, Kellar KJ (1992)
receptors in the ventral tegmental area. Psychopharmacology A subtype of nicotinic cholinergic receptor in rat brain is
149:107–114 composed of alpha 4 and beta 2 subunits and is up-regulated
Costall B, Jones BJ, Kelly ME, Naylor RJ, Tomkins DM (1989) by chronic nicotine treatment. Mol Pharmacol 41:31–37
Exploration of mice in a black and white test box: validation Flores CM, Davila-Garcia MI, Ulrich YM, Kellar KJ (1997)
as a model of anxiety. Pharmacol Biochem Behav 32:777–785 Differential regulation of neuronal nicotinic receptor binding
Court JA, Lloyd S, Thomas N, Piggott MA, Marshall EF, sites following chronic nicotine administration. J Neurochem
Morris CM, Lamb H, Perry RH, Johnson M, Perry EK (1998) 69:2216–2219
Dopamine and nicotinic receptor binding and the levels of Franceschini D, Orr-Urtreger A, Yu W, Mackey LY, Bond RA,
dopamine and homovanillic acid in human brain related to Armstrong D, Patrick JW, Beaudet AL, De Biasi M (2000)
tobacco use. Neuroscience 87:63–78 Altered baroreflex response in alpha7 deficient mice. Behav
Coutcher JB, Cawley G, Wecker L (1992) Dietary choline supple- Brain Res 113:3–10
mentation increases the density of nicotine binding sites in rat Fudala PJ, Iwamoto ET (1986) Further studies on nicotine-
brain. J Pharmacol Exp Ther 262:1128–1132 induced conditioned place preference in the rat. Pharmacol
Cox BM, Goldstein A, Nelson WT (1984) Nicotine self-adminis- Biochem Behav 25:1041–1049
tration in rats. Br J Pharmacol 83:49–55 Fudala PJ, Teoh KW, Iwamoto ET (1985) Pharmacologic character-
Damsma G, Day J, Fibiger HC (1989) Lack of tolerance to nicotine- ization of nicotine-induced conditioned place preference.
induced dopamine release in the nucleus accumbens. Eur Pharmacol Biochem Behav 22:237–241
J Pharmacol 168:363–368 Fung YK, Schmid MJ, Anderson TM, Lau YS (1996) Effects of
Deutch AY, Holliday J, Roth RH, Chun LL, Hawrot E (1987) nicotine withdrawal on central dopaminergic systems. Pharmacol
Immunohistochemical localization of a neuronal nicotinic Biochem Behav 53:635–640
115
Ghosheh OA, Dwoskin LP, Miller DK, Crooks PA (2001) Imperato A, Mulas A, Di Chiara G (1986) Nicotine preferentially
Accumulation of nicotine and its metabolites in rat brain stimulates dopamine release in the limbic system of freely
after intermittent or continuous peripheral administration of moving rats. Eur J Pharmacol 132:337–338
[2′-(14)C]nicotine. Drug Metab Dispos 29:645–651 Irvine EE, Cheeta S, File SE (1999) Time-course of changes in the
Glick SD, Visker KE, Maisonneuve IM (1996) An oral self- social interaction test of anxiety following acute and chronic
administration model of nicotine preference in rats: effects of administration of nicotine. Behav Pharmacol 10:691–697
mecamylamine. Psychopharmacology 128:426–431 Irvine EE, Bagnalasta M, Marcon C, Motta C, Tessari M, File SE,
Goldberg SR, Spealman RD (1982) Maintenance and suppression Chiamulera C (2001a) Nicotine self-administration and with-
of behavior by intravenous nicotine injections in squirrel drawal: modulation of anxiety in the social interaction test in
monkeys. Fed Proc 41:216–220 rats. Psychopharmacology 153:315–320
Goldberg SR, Spealman RD, Goldberg DM (1981) Persistent Irvine EE, Cheeta S, File SE (2001b) Tolerance to nicotine’s
behavior at high rates maintained by intravenous self- effects in the elevated plus-maze and increased anxiety during
administration of nicotine. Science 214:573–575 withdrawal. Pharmacol Biochem Behav 68:319–325
Goldberg SR, Spealman RD, Risner ME, Henningfield JE (1983) Ise Y, Narita M, Nagase H, Suzuki T (2000) Modulation of
Control of behavior by intravenous nicotine injections in opioidergic system on mecamylamine-precipitated nicotine-
laboratory animals. Pharmacol Biochem Behav 19:1011–1020 withdrawal aversion in rats. Psychopharmacology 151:49–54
Gorelick DA, Rose J, Jarvik ME (1988) Effect of naloxone on Isola R, Vogelsberg V, Wemlinger TA, Neff NH, Hadjiconstantinou
cigarette smoking. J Subst Abuse 1:153–159 M (1999) Nicotine abstinence in the mouse. Brain Res 850:
Gourlay SG, Benowitz (1997) Arteriovenous differences in plasma 189–196
concentration of nicotine and catecholamines and related Ivanova S, Greenshaw AJ (1997) Nicotine-induced decreases
cardiovascular effects after smoking, nicotine nasal spray and in VTA electrical self-stimulation thresholds: blockade by
intravenous nicotine. Clin Pharmacol Ther 62:453–463 haloperidol and mecamylamine but not scopolamine or
Green TA, Phillips SB, Crooks PA, Dwoskin LP, Bardo MT (2000) ondansetron. Psychopharmacology 134:187–192
Nornicotine pretreatment decreases intravenous nicotine self- Jorenby DE, Steinpreis RE, Sherman JE, Baker TB (1990) Aversion
administration in rats. Psychopharmacology 152:289–294 instead of preference learning indicated by nicotine place
Grenhoff J, Aston-Jones G, Svensson TH (1986) Nicotinic effects conditioning in rats. Psychopharmacology 101:533–538
on the firing pattern of midbrain dopamine neurons. Acta Kassiou M, Eberl S, Meikle SR, Birrell A, Constable C,
Physiol Scand 128:351–358 Fulham MJ, Wong DF, Musachio JL (2001) In vivo imaging of
Harris CM, Emmett-Oglesby MW, Robinson NG, Lal H (1986) nicotinic receptor upregulation following chronic (–)-nicotine
Withdrawal from chronic nicotine substitutes partially for the treatment in baboon using SPECT. Nucl Med Biol 28:165–175
interoceptive stimulus produced by pentylenetetrazol (PTZ). Koylu E, Demirgoren S, London ED, Pogun S (1997) Sex difference
Psychopharmacology 90:85–89 in up-regulation of nicotinic acetylcholine receptors in rat
Heidbreder CA, Babovic-Vuksanovic D, Shoaib M, Shippenberg TS brain. Life Sci 61:PL185–190
(1995) Development of behavioral sensitization to cocaine: Ksir C, Hakan R, Hall DP Jr, Kellar KJ (1985) Exposure to nicotine
influence of kappa opioid receptor agonists. J Pharmacol Exp enhances the behavioral stimulant effect of nicotine and
Ther 275:150–163 increases binding of [3H]acetylcholine to nicotinic receptors.
Heidbreder C, Shoaib M, Shippenberg TS (1996) Differential role Neuropharmacology 24:527–531
of delta-opioid receptors in the development and expression of Ksir C, Hakan RL, Kellar KJ (1987) Chronic nicotine and locomotor
behavioral sensitization to cocaine. Eur J Pharmacol 298: activity: influences of exposure dose and test dose. Psycho-
207–216 pharmacology 92:25–29
Helton DR, Modlin DL, Tizzano JP, Rasmussen K (1993) Nicotine Kyerematen GA, Owens GF, Chattopadhyay B, deBethizy JD,
withdrawal: a behavioral assessment using schedule controlled Vesell ES (1988) Sexual dimorphism of nicotine metabolism
responding, locomotor activity, and sensorimotor reactivity. and distribution in the rat. Studies in vivo and in vitro. Drug
Psychopharmacology 113:205–210 Metab Dispos 16:823–828
Helton DR, Tizzano JP, Monn JA, Schoepp DD, Kallman MJ Lapchak PA, Araujo DM, Quirion R, Collier B (1989) Effect of
(1997) LY354740: a metabotropic glutamate receptor agonist chronic nicotine treatment on nicotinic autoreceptor function
which ameliorates symptoms of nicotine withdrawal in rats. and N-[3H]methylcarbamylcholine binding sites in the rat
Neuropharmacology 36:1511–1516 brain. J Neurochem 52:483–491
Henningfield JE, Goldberg SR (1983) Nicotine as a reinforcer in Lê AD, Corrigall WA, Wachtus J, Harding S, Juzytsch W, Li TK
human subjects and laboratory animals. Pharmacol Biochem (2000) Involvement of nicotinic receptors in alcohol self-
Behav 19:989–992 administration. Alcohol Clin Exp Res 24:155–163
Hildebrand BE, Nomikos GG, Bondjers C, Nisell M, Svensson TH Le Novere N, Zoli M, Changeux JP (1996) Neuronal nicotinic
(1997) Behavioral manifestations of the nicotine abstinence receptor alpha 6 subunit mRNA is selectively concentrated in
syndrome in the rat: peripheral versus central mechanisms. catecholaminergic nuclei of the rat brain. Eur J Neurosci 8:
Psychopharmacology 129:348–356 2428–2439
Hildebrand BE, Panagis G, Svensson TH, Nomikos GG (1999) Levin ED, Morgan MM, Galvez C, Ellison GD (1987) Chronic
Behavioral and biochemical manifestations of mecamylamine- nicotine and withdrawal effects on body weight and food
precipitated nicotine withdrawal in the rat: role of nicotinic and water consumption in female rats. Physiol Behav 39:
receptors in the ventral tegmental area. Neuropsychopharma- 441–444
cology 21:560–574 Lichtensteiger W, Hefti F, Felix D, Huwyler T, Melamed E,
Hollt V, Horn G (1992) Effect of nicotine on mRNA levels encoding Schlumpf M (1982) Stimulation of nigrostriatal dopamine
opioid peptides, vasopressin and alpha 3 nicotinic receptor neurones by nicotine. Neuropharmacology 21:963–968
subunit in the rat. Clin Invest 70:224–231 London ED, Waller SB, Wamsley JK (1985) Autoradiographic
Houdi AA, Dasgupta R, Kindy MS (1998) Effect of nicotine use localization of [3H]nicotine binding sites in the rat brain.
and withdrawal on brain preproenkephalin A mRNA. Brain Neurosci Lett 53:179–184
Res 799:257–263 Lukas RJ, Changeux JP, Le Novere N, Albuquerque EX,
Huston-Lyons D, Kornetsky C (1992) Effects of nicotine on the Balfour DJ, Berg DK, Bertrand D, Chiappinelli VA, Clarke PB,
threshold for rewarding brain stimulation in rats. Pharmacol Collins AC, Dani JA, Grady SR, Kellar KJ, Lindstrom JM,
Biochem Behav 41:755–759 Marks MJ, Quik M, Taylor PW, Wonnacott S (1999) Interna-
Huston-Lyons D, Sarkar M, Kornetsky C (1993) Nicotine and brain- tional Union of Pharmacology. XX. Current status of the
stimulation reward: interactions with morphine, amphetamine nomenclature for nicotinic acetylcholine receptors and their
and pimozide. Pharmacol Biochem Behav 46:453–457 subunits. Pharmacol Rev 51:397–401
116
Lunell E, Molander L, Ekberg K, Wahren J (2000) Site of nicotine “binge” nicotine administration. Drug Alcohol Depend 60
absorption from a vapour inhaler-comparison with cigarette [Suppl 1]:S142
smoking. Eur J Clin Pharmacol 55:737–741 McEvoy JP, Freudenrich O, Levin ED, Rose JE (1995) Haloperidol
Lynch WJ, Carroll ME (1999) Regulation of intravenously self- increases smoking in patients with schizophrenia. Psycho-
administered nicotine in rats. Exp Clin Psychopharmacol 7: pharmacology 119:124–126
198–207 Mereu G, Yoon KW, Boi V, Gessa GL, Naes L, Westfall TC
Maehler R, Dadmarz M, Vogel WH (2000) Determinants of the (1987) Preferential stimulation of ventral tegmental area dopa-
voluntary consumption of nicotine by rats. Neuropsychobiology minergic neurons by nicotine. Eur J Pharmacol 141:395–399
41:200–204 Mifsud JC, Hernandez L, Hoebel BG (1989) Nicotine infused into
Maisonneuve IM, Glick SD (1999) (+/–) Cyclazocine blocks the the nucleus accumbens increases synaptic dopamine as
dopamine response to nicotine. Neuroreport 10:693–696 measured by in vivo microdialysis. Brain Res 478:365–367
Malin DH, Lake JR, Newlin-Maultsby P, Roberts LK, Lanier JG, Miller RP, Rotenberg KS, Adir J (1977) Effect of dose on the
Carter VA, Cunningham JS, Wilson OB (1992) Rodent model pharmacokinetics of intravenous nicotine in the rat. Drug
of nicotine abstinence syndrome. Pharmacol Biochem Behav Metab Dispos 5:436–443
43:779–784 Miyata H, Ando K, Yanagita T (1999) Medial prefrontal cortex is
Malin DH, Lake JR, Carter VA, Cunningham JS, Wilson OB involved in the discriminative stimulus effects of nicotine in
(1993) Naloxone precipitates nicotine abstinence syndrome in rats. Psychopharmacology 145:234–236
the rat. Psychopharmacology 112:339–342 Museo E, Wise RA (1994) Place preference conditioning with
Malin DH, Lake JR, Carter VA, Cunningham JS, Hebert KM, ventral tegmental injections of cytisine. Life Sci 55:1179–1186
Conrad DL, Wilson OB (1994) The nicotinic antagonist Nadal R, Samson HH (1999) Operant ethanol self-administration
mecamylamine precipitates nicotine abstinence syndrome in after nicotine treatment and withdrawal. Alcohol 17:139–147
the rat. Psychopharmacology 115:180–184 NIDA (1998) Nicotine addiction. NIDA research report series,
Malin DH, Lake JR, Short PE, Blossman JB, Lawless BA, NIH publication no 98-4342
Schopen CK, Sailer EE, Burgess K, Wilson OB (1996) Nicotine Nisell M, Nomikos GG, Svensson TH (1994) Systemic nicotine-
abstinence syndrome precipitated by an analog of neuro- induced dopamine release in the rat nucleus accumbens is
peptide FF. Pharmacol Biochem Behav 54:581–585 regulated by nicotinic receptors in the ventral tegmental area.
Malin DH, Lake JR, Upchurch TP, Shenoi M, Rajan N, Synapse 16:36–44
Schweinle WE (1998) Nicotine abstinence syndrome precipi- Nisell M, Nomikos GG, Hertel P, Panagis G, Svensson TH (1996)
tated by the competitive nicotinic antagonist dihydro-beta- Condition-independent sensitization of locomotor stimulation
erythroidine. Pharmacol Biochem Behav 60:609–613 and mesocortical dopamine release following chronic nicotine
Mansbach RS, Chambers LK, Rovetti CC (2000) Effects of the treatment in the rat. Synapse 22:369–381
competitive nicotinic antagonist, erysodine on behavior Nisell M, Marcus M, Nomikos GG, Svensson TH (1997) Differential
occasioned or maintained by nicotine: comparison with meca- effects of acute and chronic nicotine on dopamine output in
mylamine. Psychopharmacology 148:234–242 the core and shell of the rat nucleus accumbens. J Neural
Markou A, Weiss F, Gold LH, Caine SB, Schulteis G, Koob GF Transm 104:1–10
(1993) Animal models of drug craving. Psychopharmacology Nordberg A, Romanelli L, Sundwall A, Bianchi C, Beani L (1989)
112:163–182 Effect of acute and subchronic nicotine treatment on cortical
Marks MJ, Burch JB, Collins AC (1983) Effects of chronic nicotine acetylcholine release and on nicotinic receptors in rats and
infusion on tolerance development and nicotinic receptors guinea-pigs. Br J Pharmacol 98:71–78
J Pharmacol Exp Ther 226:817–825 Orr-Urtreger A, Goldner FM, Saeki M, Lorenzo I, Goldberg L, De
Marks MJ, Pauly JR, Gross SD, Deneris ES, Hermans-Borgmeyer I, Biasi M, Dani JA, Patrick JW, Beaudet AL (1997) Mice deficient
Heinemann SF, Collins AC (1992) Nicotine binding and in the alpha7 neuronal nicotinic acetylcholine receptor lack
nicotinic receptor subunit RNA after chronic nicotine treatment. alpha-bungarotoxin binding sites and hippocampal fast nico-
J Neurosci 12:2765–2784 tinic currents. J Neurosci 17:9165–9171
Marks MJ, Robinson SF, Collins AC (1996) Nicotinic agonists Ouagazzal A-M, Kenny PJ, File SE (1999) Modulation of behaviour
differ in activation and desensitization of 86Rb+ efflux from on trials 1 and 2 in the elevated plus-maze test of anxiety after
mouse thalamic synaptosomes. J Pharmacol Exp Ther 277: systemic and hippocampal administration of nicotine. Psycho-
1383–1396 pharmacology 144:54–60
Marshall DL, Redfern PH, Wonnacott S (1997) Presynaptic nico- Paylor R, Nguyen M, Crawley JN, Patrick J, Beaudet A,
tinic modulation of dopamine release in the three ascending Orr-Urtreger A (1998) Alpha7 nicotinic receptor subunits are
pathways studied by in vivo microdialysis: comparison of not necessary for hippocampal-dependent learning or sensori-
naive and chronic nicotine-treated rats. J Neurochem 68: motor gating: a behavioral characterization of alpha7-deficient
1511–1519 mice. Learning and Memory 5:302-316
Marubio LM, Changeux J (2000) Nicotinic acetylcholine receptor Panagis G, Nisell M, Nomikos GG, Chergui K, Svensson TH
knockout mice as animal models for studying receptor function. (1996) Nicotine injections into the ventral tegmental area
Eur J Pharmacol 393:113–121 increase locomotion and Fos-like immunoreactivity in the
Marubio LM, Mar Arroyo-Jimenez M, Cordero-Erausquin M, nucleus accumbens of the rat. Brain Res 730:133–142
Lena C, Le Novere N, de Kerchove d’Exaerde A, Huchet M, Paterson D, Nordberg A (2000) Neuronal nicotinic receptors in the
Damaj MI, Changeux JP (1999) Reduced antinociception in human brain. Prog Neurobiol 61:75–111
mice lacking neuronal nicotinic receptor subunits. Nature 398: Pauly JR, Marks MJ, Gross SD, Collins AC (1991) An autoradio-
805–810 graphic analysis of cholinergic receptors in mouse brain
Mathieu AM, Caboche J, Besson MJ (1996) Distribution of pre- after chronic nicotine treatment. J Pharmacol Exp Ther 258:
proenkephalin, preprotachykinin A, and preprodynorphin 1127–1136
mRNAs in the rat nucleus accumbens: effect of repeated Pauly JR, Marks MJ, Robinson SF, van de Kamp JL, Collins AC
administration of nicotine. Synapse 23:94–106 (1996) Chronic nicotine and mecamylamine treatment increase
Mathieu-Kia AM, Besson MJ (1998) Repeated administration of brain nicotinic receptor binding without changing alpha 4 or
cocaine, nicotine and ethanol: effects on preprodynorphin, beta 2 mRNA levels. J Pharmacol Exp Ther 278:361–369
preprotachykinin A and preproenkephalin mRNA expression Perez-Stable EJ, Herrera B, Jacob P, III, Benowitz NL (1998)
in the dorsal and the ventral striatum of the rat. Brain Res Mol Nicotine metabolism and intake in black and white smokers.
Brain Res 54:141–151 JAMA 280:152–156
Mathieu-Kia AM, Schlussman SD, Yuferov V, Ho A, Kreek MJ Perkins KA, Sexton JE, Reynolds WA, Grobe JE, Fonte C,
(2000) Opioid peptide mRNA levels in the rat brain after Stiller RL (1994) Comparison of acute subjective and heart
117
rate effects of nicotine intake via tobacco smoking versus Romanelli L, Ohman B, Adem A, Nordberg A (1988) Subchronic
nasal spray. Pharmacol Biochem Behav 47:295–299 treatment of rats with nicotine: interconversion of nicotinic
Perkins KA, Donny E, Caggiula AR (1999) Sex differences in receptor subtypes in brain. Eur J Pharmacol 148:289–291
nicotine effects and self-administration: review of human and Rose JE, Corrigall WA (1997) Nicotine self-administration in
animal evidence. Nicotine Tob Res 1:301–315 animals and humans: similarities and differences. Psychopharma-
Picciotto MR, Zoli M, Rimondini R, Lena C, Marubio LM, cology 130:28–40
Pich EM, Fuxe K, Changeux JP (1998) Acetylcholine receptors Rose JE, Behm FM, Westman EC, Coleman RE (1999) Arterial
containing the beta2 subunit are involved in the reinforcing nicotine kinetics during cigarette smoking and intravenous
properties of nicotine. Nature 391:173–177 nicotine administration: implications for addiction. Drug Alcohol
Picciotto MR, Caldarone BJ, King SL, Zachariou V (2000) Nicotinic Depend 56:99–107
receptors in the brain. Links between molecular biology and Rosecrans JA, Karan LD (1993) Neurobehavioral mechanisms of
behavior. Neuropsychopharmacology 22:451–465 nicotine action: role in the initiation and maintenance of tobacco
Pierzchala K, Houdi AA, Van Loon GR (1987) Nicotine-induced dependence. J Subst Abuse Treat 10:161–170
alterations in brain regional concentrations of native and Rosecrans JA, Hendry JS, Hong JS (1985) Biphasic effects of
cryptic Met- and Leu-enkephalin. Peptides 8:1035–1043 chronic nicotine treatment on hypothalamic immunoreactive
Pietila K, Ahtee L (2000) Chronic nicotine administration in the beta-endorphin in the mouse. Pharmacol Biochem Behav 23:
drinking water affects the striatal dopamine in mice. Pharmacol 141–143
Biochem Behav 66:95–103 Ross SA, Wong JY, Clifford JJ, Kinsella A, Massalas JS,
Pietila K, Lahde T, Attila M, Ahtee L, Nordberg A (1998) Regula- Horne MK, Scheffer IE, Kola I, Waddington JL, Berkovic SF,
tion of nicotinic receptors in the brain of mice withdrawn from Drago J (2000) Phenotypic characterization of an alpha 4
chronic oral nicotine treatment. Naunyn Schmiedebergs Arch neuronal nicotinic acetylcholine receptor subunit knock-out
Pharmacol 357:176–182 mouse. J Neurosci 20:6431–6441
Plowchalk DR, Andersen ME, deBethizy JD (1992) A physio- Rossi FM, Pizzorusso T, Porciatti V, Marubio LM, Maffei L,
logically based pharmacokinetic model for nicotine disposition Changeux JP (2001) Requirement of the nicotinic acetylcholine
in the Sprague-Dawley rat. Toxicol Appl Pharmacol 116: receptor beta 2 subunit for the anatomical and functional
177–188 development of the visual system. Proc Natl Acad Sci USA
Pomerleau OF (1998) Endogenous opioids and smoking: a review 98:6453-6458
of progress and problems. Psychoneuroendocrinology 23: Rotenberg KS, Miller RP, Adir J (1980) Pharmacokinetics of nicotine
115–130 in rats after single-cigarette smoke inhalation. J Pharm Sci
Pontieri FE, Tanda G, Orzi F, Di Chiara G (1996) Effects of 69:1087–1090
nicotine on the nucleus accumbens and similarity to those of Rowell PP, Li M (1997) Dose-response relationship for nicotine-
addictive drugs. Nature 382:255–257 induced up-regulation of rat brain nicotinic receptors. J Neuro-
Quinn DI, Wodak A, Day RO (1997) Pharmacokinetic and chem 68:1982–1989
pharmacodynamic principles of illicit drug use and treatment Ryan RE, Loiacono RE (2001) Nicotine regulates alpha7 nicotinic
of illicit drug users. Clin Pharmacokinet 33:344–400 receptor subunit mRNA: implications for nicotine dependence.
Rasmussen DD (1998) Effects of chronic nicotine treatment Neuroreport 12:569–572
and withdrawal on hypothalamic proopiomelanocortin gene Sastry BV, Chance MB, Singh G, Horn JL, Janson VE (1995)
expression and neuroendocrine regulation. Psychoneuroendo- Distribution and retention of nicotine and its metabolite,
crinology 23:245–259 cotinine, in the rat as a function of time. Pharmacology 50:
Rasmussen K, Czachura JF (1995) Nicotine withdrawal leads to 128–136
increased firing rates of midbrain dopamine neurons. Neuro- Schneider NG, Lunell E, Olmstead RE, Fagerstrom KO (1996)
report 7:329–332 Clinical pharmacokinetics of nasal nicotine delivery. A review
Rasmussen T, Swedberg MD (1998) Reinforcing effects of and comparison to other nicotine systems. Clin Pharmacokinet
nicotinic compounds: intravenous self-administration in drug- 31:65–80
naive mice. Pharmacol Biochem Behav 60:567–573 Schwartz RD, Kellar KJ (1983) Nicotinic cholinergic receptor
Rasmussen K, Kallman MJ, Helton DR (1997) Serotonin-1A binding sites in the brain: regulation in vivo. Science 220:
antagonists attenuate the effects of nicotine withdrawal on the 214–216
auditory startle response. Synapse 27:145–152 Schwartz RD, Kellar KJ (1985) In vivo regulation of [3H]acetyl-
Reavill C, Stolerman IP (1990) Locomotor activity in rats after choline recognition sites in brain by nicotinic cholinergic drugs.
administration of nicotinic agonists intracerebrally. Br J Pharma- J Neurochem 45:427–433
col 99:273–278 Semba J, Mataki C, Yamada S, Nankai M, Toru M (1998) Antide-
Reid MS, Ho LB, Berger SP (1996) Effects of environmental pressant-like effects of chronic nicotine on learned helplessness
conditioning on the development of nicotine sensitization: paradigm in rats. Biol Psychiatry 43:389–391
behavioral and neurochemical analysis. Psychopharmacology Shaham Y, Adamson LK, Grocki S, Corrigall WA (1997) Reinstate-
126:301–310 ment and spontaneous recovery of nicotine seeking in rats.
Risinger FO, Oakes RA (1995) Nicotine-induced conditioned Psychopharmacology 130:396–403
place preference and conditioned place aversion in mice. Shoaib M, Stolerman IP (1999) Plasma nicotine and cotinine
Pharmacol Biochem Behav 51:457–461 levels following intravenous nicotine self-administration in
Risner ME, Goldberg SR (1983) A comparison of nicotine and rats. Psychopharmacology 143:318–321
cocaine self-administration in the dog: fixed-ratio and progres- Shoaib M, Stolerman IP, Kumar RC (1994) Nicotine-induced
sive-ratio schedules of intravenous drug infusion. J Pharmacol place preferences following prior nicotine exposure in rats.
Exp Ther 224:319–326 Psychopharmacology 113:445–452
Robinson SF, Marks MJ, Collins AC (1996) Inbred mouse strains Shoaib M, Schindler CW, Goldberg SR (1997) Nicotine self-
vary in oral self-selection of nicotine. Psychopharmacology administration in rats: strain and nicotine pre-exposure effects
124:332–339 on acquisition. Psychopharmacology 129:35–43
Rogers WR, Wilbur RL, Bass RL, III, Johnson DE (1985) Effects Shoaib M, Swanner LS, Yasar S, Goldberg SR (1999) Chronic
of cigarette nicotine content on smoking behavior of baboons. caffeine exposure potentiates nicotine self-administration in
Addict Behav 10:225–233 rats. Psychopharmacology 142:327–333
Rogers WR, Carey KD, McMahan CA, Montiel MM, Mott GE, Singer G, Wallace M, Hall R (1982) Effects of dopaminergic
Wigodsky HS, McGill HC Jr (1988) Cigarette smoking, nucleus accumbens lesions on the acquisition of schedule
dietary hyperlipidemia, and experimental atherosclerosis in the induced self injection of nicotine in the rat. Pharmacol Biochem
baboon. Exp Mol Pathol 48:135–151 Behav 17:579–581
118
Sorenson EM, Shiroyama T, Kitai ST (1998) Postsynaptic nicotinic Vetter DE, Liberman MC, Mann J, Barhanin J, Boulter J,
receptors on dopaminergic neurons in the substantia nigra pars Brown MC, Saffiote-Kolman J, Heinemann SF, Elgoyhen AB
compacta of the rat. Neuroscience 87:659–673 (1999) Role of alpha9 nicotinic ACh receptor subunits in the
Stolerman IP (1989) Discriminative stimulus effects of nicotine development and function of cochlear efferent innervation.
in rats trained under different schedules of reinforcement. Neuron 23:93–103
Psychopharmacology 97:131–138 Wada E, Wada K, Boulter J, Deneris E, Heinemann S, Patrick J,
Stolerman IP (1999) Inter-species consistency in the behavioural Swanson LW (1989) Distribution of alpha 2, alpha 3, alpha 4,
pharmacology of nicotine dependence. Behav Pharmacol 10: and beta 2 neuronal nicotinic receptor subunit mRNAs in the
559–580 central nervous system: a hybridization histochemical study in
Stolerman IP, Naylor C, Elmer GI, Goldberg GR (1999) Discrimi- the rat. J Comp Neurol 284:314–335
nation and self-administration of nicotine by inbred strains of Wall A, Gong ZH, Johnson AE, Meyerson B, Zhang X (2000)
mice. Psychopharmacology 141:297–306 Cross-tolerance in drug response and differential changes in
Summers KL, Giacobini E (1995) Effects of local and repeated central nicotinic and N-methyl-D-aspartate receptor binding
systemic administration of (–)nicotine on extracellular levels following chronic treatment with either (+)- or (–)-nicotine.
of acetylcholine, norepinephrine, dopamine, and serotonin in Psychopharmacology 148:186–195
rat cortex. Neurochem Res 20:753–759 Watkins SS, Koob GF, Markou A (2000a) Neural mechanisms
Sutherland G, Stapleton JA, Russell MA, Feyerabend C (1995) underlying nicotine addiction: acute positive reinforcement
Naltrexone, smoking behaviour and cigarette withdrawal. and withdrawal. Nicotine Tob Res 2:19–37
Psychopharmacology 120:418–425 Watkins SS, Stinus L, Koob GF, Markou A (2000b) Reward and
Suto N, Austin JD, Vezina P (2001) Locomotor response to somatic changes during precipitated nicotine withdrawal in
novelty predicts a rat’s propensity to self-administer nicotine. rats: centrally and peripherally mediated effects. J Pharmacol
Psychopharmacology 158:175–180 Exp Ther 292:1053–1064
Swanson LW, Simmons DM, Whiting PJ, Lindstrom J (1987) Wewers ME, Dhatt RK, Snively TA, Tejwani GA (1999) The
Immunohistochemical localization of neuronal nicotinic effect of chronic administration of nicotine on antinociception,
receptors in the rodent central nervous system. J Neurosci 7: opioid receptor binding and met-enkelphalin levels in rats.
3334–3342 Brain Res 822:107–113
Tanda G, Goldberg SR (2000) Alteration of the behavioral effects Wise RA (1996) Addictive drugs and brain stimulation reward.
of nicotine by chronic caffeine exposure. Pharmacol Biochem Annu Rev Neurosci 19:319–340
Behav 66:47–64 Wonnacott S (1990) The paradox of nicotinic acetylcholine receptor
Tessari M, Valerio E, Chiamulera C, Beardsley PM (1995) upregulation by nicotine. Trends Pharmacol Sci 11:216–219
Nicotine reinforcement in rats with histories of cocaine self- Xu W, Gelber S, Orr-Urtreger A, Armstrong D, Lewis RA,
administration. Psychopharmacology 121:282–283 Ou CN, Patrick J, Role L, De Biasi M, Beaudet AL (1999a)
Tizabi Y, Overstreet DH, Rezvani AH, Louis VA, Clark E, Megacystis, mydriasis, and ion channel defect in mice lacking
Janowsky DS, Kling MA (1999) Antidepressant effects of the alpha3 neuronal nicotinic acetylcholine receptor. Proc Natl
nicotine in an animal model of depression. Psychopharmacology Acad Sci USA 96:5746–5751
142:193–199 Xu W, Orr-Urtreger A, Nigro F, Gelber S, Sutcliffe CB,
Todte K, Tselis N, Dadmarz M, Golden G, Ferraro T, Berrettini WH, Armstrong D, Patrick JW, Role LW, Beaudet AL, De Biasi M
Vogel WH (2001) Effects of strain, behavior and age on the (1999b) Multiorgan autonomic dysfunction in mice lacking
self-administration of ethanol, nicotine, cocaine and morphine the beta2 and the beta4 subunits of neuronal nicotinic acetyl-
by two rat strains. Neuropsychobiology 44:150–155 choline receptors. J Neurosci 19:9298–9305
Toth E, Sershen H, Hashim A, Vizi ES, Lajtha A (1992) Effect of Yates SL, Bencherif M, Fluhler EN, Lippiello PM (1995) Up-
nicotine on extracellular levels of neurotransmitters assessed regulation of nicotinic acetylcholine receptors following chronic
by microdialysis in various brain regions: role of glutamic exposure of rats to mainstream cigarette smoke or alpha 4 beta
acid. Neurochem Res 17:265–271 2 receptors to nicotine. Biochem Pharmacol 50:2001–2008
Trauth JA, Seidler FJ, McCook EC, Slotkin TA (1999) Adolescent Yu ZJ, Morgan DG, Wecker L (1996) Distribution of three nicotinic
nicotine exposure causes persistent upregulation of nicotinic receptor alpha 4 mRNA transcripts in rat brain: selective regula-
cholinergic receptors in rat brain regions. Brain Res 851:9–19 tion by nicotine administration. J Neurochem 66:1326–1329
Ulrich YM, Hargreaves KM, Flores CM (1997) A comparison of Zachariou V, Caldarone BJ, Weathers-Lowin A, George TP,
multiple injections versus continuous infusion of nicotine for Elsworth JD, Roth RH, Changeux J, Picciotto MR (2001)
producing up-regulation of neuronal [3H]-epibatidine binding Nicotine receptor inactivation decreases sensitivity to cocaine.
sites. Neuropharmacology 36:1119–1125 Neuropsychopharmacology 24:576–589
Valentine JD, Hokanson JS, Matta SG, Sharp BM (1997) Self- Zhang X, Gong ZH, Nordberg A (1994) Effects of chronic
administration in rats allowed unlimited access to nicotine. treatment with (+)- and (–)-nicotine on nicotinic acetylcholine
Psychopharmacology 133:300–304 receptors and N-methyl-D-aspartate receptors in rat brain.
Van Loon GR, Pierzchala K, Houdi AA (1991) Nicotine- Brain Res 644:32–39
induced alterations in peripheral tissue concentrations of Zoli M, Lena C, Picciotto MR, Changeux JP (1998) Identification
native and cryptic Met- and Leu-enkephalin. Neuropeptides of four classes of brain nicotinic receptors using beta2 mutant
19:35–41 mice. J Neurosci 18:4461–4472

View publication stats

You might also like