National Academies Press: OpenBook
« Previous: 8 Principal Conclusions
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 243
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 244
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 245
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 246
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 247
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 248
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 249
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 250
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 251
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 252
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 253
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 254
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 255
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 256
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 257
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 258
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 259
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 260
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 261
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 262
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 263
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 264
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 265
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 266
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 267
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 268
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 269
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 270
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 271
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 272
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 273
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 274
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 275
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 276
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 277
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 278
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 279
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 280
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 281
Suggested Citation:"9 Nicotine Pharmacology." Institute of Medicine. 2001. Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction. Washington, DC: The National Academies Press. doi: 10.17226/10029.
×
Page 282

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

9 Nicotine Pharmacology BASIC AND HUMAN PHARMACOLOGY t is some 550 years since the eponymous Jean Nicot sent tobacco and I seeds from Portugal to Paris, passing Nicotiana tabacum from the Americas to Northern Europe by way of the Iberian peninsula. Nico- tine itself was subsequently extracted and synthesized, culminating in the identification of the spatial orientation of the natural (S) isomer in the late 1970s (Domino, 1999). Up to 10% of the nicotine in tobacco smoke is the (R) isomer, probably arising from racemization during combustion (Benowitz, 1986). Nicotine has gained particular prominence as the addic- tive constituent of most tobacco based products and, to a lesser extent, as an effective insecticide. Nicotine, 3-(1-methyl-2-pyrrolidinyl) pyridine, has a molecular weight of 162.23 and is a volatile, colorless base (pKa=8.5) that turns brown and acquires the typical odor of tobacco on exposure to light. Roughly 69% of its pyrrolidine nitrogen is ionized (positively charged) at pH 7.4 and 37°C, whereas its pyridine nitrogen is un-ionized. This feature of nicotine renders its absorption and renal excretion highly pH dependent, because uncharged lipophilic bases pass easily over lipoprotein mem- branes and charged organic bases do not. For example, nicotine is prima- rily ionized at the pH (5.5) of smoke from the flue-cured tobaccos in most American cigarettes, and buccal absorption is minimal (Gorrod and Wahren, 1993). By contrast, smoke from air-cured tobaccos in pipes, cigars, and many European cigarettes is less acidic and is well absorbed through 243

244 CLEARING THE SMOKE the mouth (Armitage et al., 1978; Gori et al., 1986). Nicotine constitutes about 95% of the total alkaloid content of commercial cigarette tobacco (Gorrod and Jenner, 1975). The mechanisms by which nicotine exerts its actions at a molecular level are complex. Dale (Dale, 1914) noticed the structural similarity be- tween nicotine and acetylcholine (Ach) and the resemblance of the effects of nicotine in vivo to those of Ach after pretreatment with the muscarinic antagonist atropine. The muscarinic effects of Ach are now recognized to be mediated via one of the five heptahelical muscarinic receptors (M1- M5). Ligation of these receptors may activate downstream signaling path- ways via their interaction with diverse G proteins. Nicotinic receptors (nAchRs), by contrast, are ligand gated ion channels (Domino, 1999; Lena and Changeux, 1998). These pentamers are comprised of various combi- nations of α, β, γ, and δ subunits. Recent studies have demonstrated that specific configurations of these subunits mediate the diverse effects of nicotine. Although this area of research is evolving, the neuronal subunits that appear to be primarily responsible for the effects of nicotine contain α3,4,7 and β2 and 4 subunits. The α4β2 subtype is particularly prevalent in the brain and may be responsible for the self-administration of nicotine. Mice deficient in the β2 subunit do not self-administer nicotine (Cordero- Erausquin et al., 2000), suggesting that this subunit in particular may be important in reinforcing the effects of nicotine. In addition, some prelimi- nary evidence suggests that the α7 subunit may play a significant role in withdrawal and sensory gating functions of schizophrenics (Adler et al., 1998; Nomikos et al., 2000; Panagis et al., 2000). Localization studies have identified nAchRs in the brain, neuromuscular junctions, autonomic gan- glia, and adrenal medulla (Gundisch, 2000). Ligation of nAchRs by nico- tine opens the channel, and the ionic influx activates signal transduction pathways, culminating in release of a number of different neurotransmit- ters, which have been related to nicotine’s pharmacodynamic effects. These include dopamine (pleasure and appetite suppression), serotonin (appetite suppression and mood modulation), epinephrine and norepi- nephrine (arousal and appetite suppression), Ach (arousal and cognitive enhancement), vasopressin (memory improvement), glutamate (improve- ment in learning), β-endorphin (mood modulation and analgesia), and δ- aminobutyric acid. Nicotine also increases nAchR expression. For ex- ample, prenatal nicotine exposure upregulates the pulmonary expression of the α7 receptor subunit and consequently affects fetal lung develop- ment in monkeys (Sekhon et al., 1999). Nicotine caused lung hypoplasia and reduced surface complexity of developing alveoli in this model. Col- lagen surrounding the large airways and vessels was increased, as was the number of type II cells and neuroendocrine cells in neuropepithelial

NICOTINE PHARMACOLOGY 245 bodies. Many animal studies have also demonstrated that nicotine ad- ministration upregulates expression of nAchRs in the brain. Similarly, ligand-binding studies have demonstrated an increase in binding sites for nicotine analogues in the cerebral cortex and hippocampus of smokers compared to nonsmokers (Perry et al., 1999), although the extent to which this may contribute to the differential central effects of nicotine observed in smokers is unknown. Dopamine is believed to be the dominant neurotransmitter in the maintenance of drug-taking behavior (DiChiara, 1999; Koob, 1992). The area of the brain that is responsible for the reinforcing effects of all drugs of abuse is the mesolimbic pathway, which contains the ventral tegmental area (VTA), nucleus accumbens, amygdala, cingulate gyrus, and frontal lobe and is rich in dopamine. The VTA and nucleus accumbens seem particularly important in nicotine’s reinforcing effects. Activation of nAchRs in the VTA and other parts of the midbrain, modulates the as- cending mesolimbic dopamine system, including the nucleus accumbens (George et al., 2000; Yu et al., 2000). Nicotine self-administration behavior is diminished by either surgical or chemical ablation of dopaminergic pathways or by treatment with dopamine antagonists (Kameda et al., 2000). Nicotine evokes an increase in dopamine levels in brain microdialysis studies (Fu et al., 2000). In addition, monoamine oxidase A and B, responsible for the metabolism of dopamine, are reduced by a compound in tobacco smoke that also results in higher levels of neuro- transmitters (Quattrocki et al., 2000). The release or inhibition of other transmitters may also play a role in nicotine addiction. They may be responsible for mood modulation, the modest enhancement of performance, and the weight-reducing effects of nicotine (Benowitz, 1999; Chiodera et al., 1990; Chowdhury et al., 1989; U.S. DHHS, 1988). Mood modulation by nicotine has been a controversial topic, since laboratory studies do not validate the smoking-induced en- hancement of mood self-reported by smokers. Furthermore, individuals experience greater positive affect when smoking after a period of absti- nence. The relief of negative affect by tobacco use may be more a function of abating withdrawal symptoms (Cinciripini et al., 1997; RCP, 2000). Finally, in addition to its traditional pre- and postsynaptic actions at syn- apses and at chemoreceptors in the carotid and aortic bodies, nicotine also evokes the release of epinephrine from the adrenal medulla and may act directly to activate ion channels distinct from nAchRs. For example, nico- tine has been shown to block directly inward rectifier potassium chan- nels, an effect of potential relevance to cardiac arrhythmogenesis (Wang et al., 2000b).

246 CLEARING THE SMOKE PHARMACOKINETICS Absorption Although buccal absorption is influenced by the pH of tobacco or tobacco smoke, tobacco smoke from all sources is rapidly absorbed from the large surface area of the small airways and the alveoli, following dissolution in the pulmonary fluid at pH 7.4 (Zevin et al., 1998). Nicotine is also readily absorbed from the skin; this property has been exploited in the use of patch delivery in nicotine replacement therapy (NRT) for ciga- rette smokers (Figure 9-1). Nicotine is well tolerated as a dermal applica- tion, even in individuals who suffer from irritant skin disorders (Benowitz, 1995). 20 Ciba-Geigy Kabi-Cygnus (N=11) (N=11) 15 15 Plasma Nicotine Concentration (ng/mL) 10 10 5 5 0 Alza Elan (N=13) (N=11) 15 15 10 10 5 5 0 0 8 16 24 32 0 8 16 24 32 Hours FIGURE 9-1 Time curves of plasma nicotine concentrations. NOTE: Time curves of plasma nicotine concentration after application of four dif- ferent transdermal nicotine delivery systems. The Ciba-Geigy Habitrol, the Alza (SmithKline Beecham) Nicoderm, and the Elan Prostep patches were worn for 24 hours, while the Kabi-Cygnus (Pharmacia) Nicotrol patch was worn for 16 hours. SOURCE: Benowitz, 1993. Reprinted from Drugs 1993; 45(2):157-170 with permis- sion. © Adis International, Inc.

NICOTINE PHARMACOLOGY 247 Following ingestion, as with chewing tobacco, the pH dependence of nicotine ionization favors its absorption by the small intestine, rather than the stomach, although a timed-release preparation, which permits colonic absorption (Green et al., 1999), has also been developed for use in ulcer- ative colitis as discussed below. Although the peak levels of nicotine at- tained after chewing tobacco may approximate those after smoking, the shape of the curve of plasma concentration versus time is quite different (Figure 9-2). Thus, after smoking a cigarette, plasma levels of nicotine rise rapidly to a peak, which is maintained transiently after the cigarette is inhaled, rather than the more gradual and sustained elevation after oral 0 Cigarettes Oral Snuff 15 Blood Nicotine Concentration (ng/mL) 10 5 0 Chewing Tobacco Nicotine Gum 15 10 5 0 –10 0 30 60 90 120 30 0 30 60 90 120 Minutes FIGURE 9-2 Blood nicotine concentrations. NOTE: Blood nicotine concentrations during and after cigarette smoking for nine minutes, oral snuff (2.5 grams), chewing tobacco (average 7.9 grams), and nicotine gum (two 2 mg pieces). Data represent average values for 10 subjects (±SEM). Horizontal bars above time axis indicate period of tobacco or nicotine gum exposure. SOURCE: Benowitz et al., 1988. Reprinted with permission from Mosby, Inc.

248 CLEARING THE SMOKE ingestion. The evoked liking, systemic response, and addiction potential of the former pattern of nicotine delivery exceed those of the latter (Benowitz et al., 1988). It takes roughly 10-15 seconds for nicotine, inhaled by puffing a cigarette, to reach the brain, and puffing is associated with a marked arterial-venous gradient of nicotine (Benowitz, 1995; Benowitz et al., 1988; Guthrie et al., 1999). This rapid central nervous system (CNS) delivery permits the smoker to adjust the nicotine dosage to a desired effect, reinforcing self-administration and facilitating the development of addiction (Benowitz, 1995). This contrasts with the slower increase and lesser increment in brain nicotine attained after transdermal delivery, which facilitates the development of tolerance (see below). The average cigarette contains 10-15 mg of nicotine and delivers, on average, roughly 1-2 mg of nicotine systemically to the smoker. However, smoking habit— puff intensity, duration, and so forth—can markedly alter nicotine bio- availability. By comparison, the systemic doses of nicotine from other delivery systems are roughly 1 mg from a 2 mg gum; 5-22 mg per day from transdermal patches; 0.5 mg per dose from one spray per nostril; 3.6 mg from 2.5 grams of snuff, held in the mouth for 30 minutes and 4.5 mg from 7.9 grams of chewing tobacco chewed for 30 minutes (Benowitz and Jacob, 1999). Following its absorption, nicotine circulates with roughly 60% in the ionized form. It is poorly (around 5%) protein bound (Benowitz et al., 1982) and widely distributed, at least in rats and rabbits, particularly in liver, lungs, and brain (Benowitz et al., 1990). Distribution and Metabolism The presence of both aromatic and aliphatic carbon and nitrogen at- oms in nicotine affords multiple sites for metabolic oxidation and subse- quent conjugation reactions (Figure 9-3). The disposition of nicotine has been reviewed in depth elsewhere (Benowitz and Jacob, 1999; Gorrod and Schepers, 1999). Briefly, roughly 80% of the metabolic inactivation of nico- tine involves oxygenation of the 5'-carbon to yield cotinine. This appears to involve an intermediate cytochrome P-450 (CYP)- derived 1',5'- imminium ion, which is further metabolized by aldehyde oxidase to yield cotinine (Brandange and Lindblom, 1979; Gorrod and Hibberd, 1982; Murphy, 1973). This iminium ion is an alkylating agent and has been speculated to have relevance to the carcinogenicity of tobacco, although this is not established (Hibberd and Gorrod, 1983). Oxidation of this radical may also yield nornicotine or 4-(3-pyridyl)-4-oxo-N-methylbutylamine. CYP2A6 and, to a lesser extent, 2B6 appear to play the predominant roles in nicotine carbon oxidation in humans (Benowitz and Jacob, 1999; Nakajima et al., 2001; Nakajima et al., 2000; Yamazaki et al., 1999). Although roughly

H H H H H N O N O N N N + CH3 CH3 O CH3 CH3 + N N N CH3 N +N + + CH3 O Nicotine CH3 Cotinine O Nicotine - 1 -N-oxide OH OH Nicotine Cotinine Glucuronide Glucuronide Isomethonium IOn Methonium Ion O2C O2C OH OH OH OH 4′ 4′ 3′ 3′ CO2 H H H 2′ H H 5′ N+ 5′ N1′ O N O 2′ N1′ CH3 CH3 CH3 N CH3 N N + N N O 3-Pyridylacelic NICOTINE Nicotine - ∆ 1′(5′) COTININE Iminium Ion Cotinine N-oxide Acid H N O H OH CH3 OH N H HO H H CO2H N′-Hydroxymethyl- N N N O O N O norcolinine H N CH3 N CH3 N N N H Trans-3′- 5-(3-Pyridyl)-tetrahydro- Nornicoline 5′-Hydroxycolinine 4-(3-Pyridyl)-butyric Hydroxycolinine N O Acid furan-2-one H N Norcotinine H O OH OH H O CO2H CO2H CO2H O N HO2C OH OH O N O HN CH3 N N N N CH3 N 4-(3-Pyridyl)-3- γ-3-Pyridyl-γ-hydroxy- γ-3-Pyridyl-γ-oxo- γ-3-Pyridyl-γ-oxo Trans-3′-Hydroxycotinine butenoic Acid butyric acid butyric acid N-methylbutyramide Glucuronide FIGURE 9-3 Nicotine metabolic pathways. SOURCE: Benowitz and Jacob, 1999. Copyright (1999) Wiley-Liss, Inc. Reprinted with permission of Wiley-Liss, Inc., a subsid- 249 iary of John Wiley & Sons. Translated by permission of John Wiley & Sons, Inc. All rights reserved.

250 CLEARING THE SMOKE 15% of a nicotine dose is excreted in human urine unchanged, all of the primary metabolites, including cotinine, are subject to further oxidation reactions. Oxidation appears to involve only the alicyclic pyrrolidine ni- trogen in biological systems (Gorrod and Schepers, 1999). Nicotine-1'-N- oxide may be reduced to nicotine in man by gut bacteria (Dajani et al., 1975). Phase two metabolites can be formed by methylation, glucuronida- tion, sulfation, or glutathione conjugation reactions with primary oxida- tion metabolites. Formation of such polar, water soluble molecules facili- tates excretion. Although great interindividual variation is noticeable, glucuronides may account for roughly 40% of the urinary nicotine me- tabolites in humans. This variability may also be apparent among ethnic groups. Thus, the most abundant phase 2 metabolite in the urine of North Americans is the N-glucuronide of cotinine, whereas in Europeans the O- glucuronide of trans-3'-hydroxycotinine predominates (Gorrod and Schepers, 1999). Similarly, the metabolism of nicotine is slower in African Americans than in Caucasians, due to both slower oxidative metabolism of nicotine to cotinine and slower N-glucuronidation (Benowitz et al., 1999; Caraballo et al., 1998; Perez-Stable et al., 1998). Asian Americans also metabolize both nicotine and cotinine more slowly than do Cauca- sians. Nicotine clearance declines with age (Molander et al., 2001). Although there is some evidence for differences between men and women in the pharmacodynamic response to nicotine (Pomerleau et al., 1991), this does not appear to reflect systematic differences in nicotine pharmaco- kinetics. Nicotine readily crosses the placental barrier, although there is no apparent conversion of nicotine to cotinine by placental tissues or microsomal fractions (Pastrakuljic et al., 1998). Although the potential for fetal toxicity must be considered in women undergoing NRT (as discussed below), this consideration usually occurs in the context of relativity. Thus, the hazard to the fetus of maternal cigarette smoking is well established (Oncken et al., 1998; Robinson et al., 2000), whereas the theoretically much smaller risk of NRT remains entirely notional. Clearance of nicotine falls with hepatic blood flow during sleep (Gries et al, 1996) and a circadian pattern in both circulating nicotine and cotinine is evident (Figure 9-4). Although the half-life of nicotine is about 2-3 hours when based on plasma levels, it approximates 11 hours when based on urinary excretion (Benowitz and Jacob, 1994, 1999), so circulating levels tend to accumulate during the day. The afternoon levels of nicotine in the plasma of smokers generally range from 10 to 50 ng/ml, whereas steady- state levels with patches range from 10 to 20 ng/ml and with the nasal spray from 5 to 15 ng/ml (Benowitz and Jacob, 1999). Sophisticated approaches to analysis not just of nicotine and cotinine, but of minor oxidative metabolites and many phase 2 metabolites, have

NICOTINE PHARMACOLOGY 251 40 30 Nicotine 20 Blood Concentrations (ng/mL) 10 0 400 360 320 Cotinine 280 240 200 0800 1200 1600 2000 2400 0400 0800 Clock Time FIGURE 9-4 Circadian blood concentrations of nicotine and cotinine during unrestricted smoking. NOTE: Data are mean ± SE for eight subjects. SOURCE: Benowitz et al., 1983. Reprinted with permission from Mosby, Inc.

252 CLEARING THE SMOKE now been established, albeit in few laboratories. These encompass sensi- tive and specific methodologies, such as tandem mass spectrometry (Byrd et al., 1994). These tools will afford a more comprehensive approach to investigating genetic (e.g., nAchR or CYP2A6 polymorphisms; McKinney et al., 2000; Nakajima et al., 1996) and environmental factors (e.g., CYP2A6 induction or repression by alcohol or other drugs; Niemela et al., 2000), which might contribute to interindividual variability in nicotine pharma- cokinetics. For example, they will include the integration of assessment of nitrosamines formed from nicotine into long-term studies of the safety of NRT. Excretion Nicotine is excreted by both glomerular filtration and tubular secre- tion. Acidification of urine greatly increases its renal clearance, which impedes tubular reabsorption by ionizing the nicotine (Benowitz and Jacob, 1999). Urinary excretion of cotinine is less influenced by pH since it is more basic. However, renal clearance of both compounds is influenced highly by urinary flow rates (Benowitz and Jacob, 1999). PHARMACODYNAMICS Cardiovascular Effects The factors that mediate the effects of nicotine are complex, con- founded as they are in the cardiovascular system by direct and reflex effects, acute effects and long-term desensitization, and secondary effects due to sympathoadrenal activation. Acute delivery of nicotine in a ciga- rette results in a transient tachycardia, cutaneous vasoconstriction, and a rise in blood pressure (Cryer et al, 1976). By contrast, desensitization of vascular or central receptors by nicotine may contribute to the lower blood pressure observed in chronic smokers (Charlton and While, 1995). The mechanisms involved in mediating the adverse effects of ciga- rette smoking and of smokeless tobacco on the cardiovascular system are poorly understood, but are thought to include induction of an adverse lipoprotein profile (Allen et al., 1994), induction of a chronic inflamma- tory response (Strandberg and Tilvis, 2000) including oxidative tissue injury (Morrow et al., 1995; Patrignani et al., 2000; Reilly et al., 1996; Traber et al., 2000), activation of platelets and other hemostatic variables (Benowitz et al., 1993; Ludviksdottir et al., 1999; Whiss et al., 2000), and impairment of endothelial function (Raitakari et al., 2000). Following the introduction of NRT, there was considerable concern about the cardio- vascular safety of this intervention, especially in patients with preexisting cardiovascular disease. However, NRT has been shown to be effective, without an apparent cardiovascular hazard, not only in the general

NICOTINE PHARMACOLOGY 253 population (Benowitz and Gourlay, 1997), but also in patients with estab- lished coronary vascular disease (Greenland et al., 1998; Joseph et al., 1996; Nitenberg and Antony, 1999). Controlled studies have demonstrated that switching from cigarette smoking to NRT is associated with amelioration of the lipoprotein and hemostatic profiles (Allen et al., 1994; Ludviksdottir et al., 1999; Winther and Fornitz, 1999) and a reduction in platelet activation (Nowak et al., 1987). Although evidence of a clinical cardiovascular hazard of NRT has yet to emerge, several observations suggest that aspects of the cardiovascular effects of nicotine merit further research. For example, cigarette smoking impairs endothelial function (Celermajer et al., 1993; Raitakari et al., 1999), which appears to be a surrogate marker of future clinical vascular disease (Dugi and Rader, 2000). Interestingly, nicotine has been reported to im- pair flow mediated brachial arterial endothelial function (Chalon et al., 2000) and bradykinin stimulated endothelial function in dorsal hand veins of volunteers (Sabha et al., 2000). On the other hand, short term adminis- tration of nicotine gum did not alter the coronary constrictor response to the cold pressor test—reflective of the effects of sympathadrenal activa- tion—in patients with established coronary artery disease (Nitenberg and Antony, 1999). Although exposure of mice deficient in apoenzyme E (Apol E) to cigarette smoke accelerates atherogenesis (Gairola and Daugherty, 1999), there have been no analogous studies of nicotine and no studies of atherosclerotic plaque progression in individuals receiving long-term NRT. Switching to NRT in the short term does not appear to correct the systemic markers of inflammation in cigarette smokers (Nilssen et al., 1996), and there are conflicting reports of the direct effects of nicotine on free radical generation in vitro (Gouaze et al., 1998; Guatura et al., 2000; Wetscher et al, 1995). Interestingly, many cigarette smokers also drink alcohol (Swahn and Hammig, 2000), and alcohol is a potent prooxidant in humans (Meagher and FitzGerald, 2000). While ethanol increases the clearance of nicotine, the acute hemodynamic effects of ethanol and nico- tine are additive (Soderpalm et al., 2000), perhaps reflecting a common mechanism. However, there are no reports of the effects of NRT on con- temporary markers of oxidative stress or a comprehensive assessment of its effects on markers of inflammation. Central Effects General Effects Nicotine is a CNS stimulant. Paradoxically, it is perceived to be relax- ing in stressful situations and to enhance gating of relevant stimuli. The smoker does not react as much as the nonsmoker to external distrac-

254 CLEARING THE SMOKE tions—hence its use by those trying to relax or concentrate. Stress in- creases the smoker’s nicotine consumption. An increase in respiratory rate and heart rate has been observed with NRT. Sleeplessness has also been reported in patients using NRT (Gourlay et al., 1999). Nicotine over- dose is remarkably difficult to achieve with NRT (Labelle and Boulay, 1999), however, it occasionally complicates the use of nicotine-containing insecticides. In these cases, central symptoms—initially tremors, nausea, vomiting, and possibly convulsions—give way to signs of central depres- sion and neuromuscular blockade (Saxena and Scheman, 1985). Memory and Cognition Although there is considerable interest in the potential effects of nico- tine on cognition (Emilien et al., 2000; Waters and Sutton, 2000), this has not been formally evaluated in individuals receiving NRT. Activation of nAchRs containing the α7 subunit results in Ach release and calcium acti- vation, and both effects have been implicated in memory formation and cognition (Kem, 2000). Recent interest has been focused by co-immuno- precipitation of the amyloid A β(1-42)-fragment and the α7nAchR from the dendtritic plaques of Alzheimer’s disease (AD) lesions (Wang et al., 2000a). The β-fragment of amyloid A binds to the receptor and prevents its activation by nicotine, potentially implicating defective nAchR activa- tion in the pathogenesis of AD. Again, although there is some evidence for a slowing of deterioration of AD in individuals who smoke (Debanne et al., 2000; Doll et al., 2000; Jarvik, 1991; Lopez-Arrieta et al., 2000; Merchant et al., 1999), along with a considerable literature relating to the use of cholinesterase inhibitors for this condition, NRT has not been for- mally evaluated in AD. Addiction The Biological Basis of Addiction. Although tobacco products contain several thousand chemicals, nicotine is considered to be the principal constituent in tobacco that leads to the persistent use of tobacco products (U.S. DHHS, 1988). However, other yet unknown constituents in tobacco may also have a role in the maintaining the use of tobacco. For example, smokers experience a reduction of monoamine oxidase (MAO) activity in the brain (Berlin et al., 1995) as a result of some constituent in smoke (Fowler et al., 1996); inhibition of MAO may result in antidepressant ac- tivity (Oxenkrug, 1999) and contribute to the high prevalence in smoking among individuals with depressive disorders. Physical addiction to nico- tine is associated with euphoriant and other psychoactive effects, the de- velopment of tolerance, and the experience of withdrawal symptoms

NICOTINE PHARMACOLOGY 255 when the tobacco product is no longer used (U.S. DHHS, 1988). In addi- tion, the rate of absorption and therefore the speed of delivery of nicotine to the brain also play a significant role in the addictive potential of nico- tine (Henningfield and Keenan, 1993). These factors contribute to the rein- forcing effects or persistent use of nicotine and also may be responsible for day-to-day regulation of nicotine levels in tobacco users. Psychoactive and reinforcing effects from nicotine are the result of the release of a number of neurotransmitters and hormones (Benowitz, 1999; U.S. DHHS, 1988; Watkins et al., 2000). This cascade of events is associ- ated with mood modulation, cognitive and motor performance enhance- ments, and weight reduction. These effects may contribute to the initia- tion and maintenance of tobacco use. Chronic administration of nicotine can lead to neuroadaptation. One of the effects of neuroadaptation is the development of tolerance. Adaptation occurs so that the brain can main- tain a state of homeostasis despite an increased release of neurotransmit- ters. This process includes receptor inactivation and desensitization and an increase or upregulation in receptor number (Benowitz, 1999). The extent of these changes could vary depending on the receptor subtype and site (Watkins et al., 2000). Tolerance may lead to individuals’ using more of the tobacco product or switching to higher nicotine-containing products. Neuroadaptation may subsequently lead to withdrawal symptoms when the tobacco user is no longer exposed to the product. Withdrawal symptoms include negative affect (e.g., irritability, frustration or anger, anxiety, dysphoric or depressed mood), restlessness, difficulty in concen- trating, insomnia, decreased heart rate, and increased appetite or weight (APA, 1994). These symptoms occur among regular users of cigarettes and smokeless tobacco (Hughes and Hatsukami, 1992). They are less pro- nounced with nicotine gum use, but this distinction blurs with prolonged use of the gum (Hughes et al., 1986b; West and Russell, 1985). Approxi- mately 49% of self-quitting smokers and 87% of tobacco cessation program attendees meet Diagnostic and Statistical Manual of Mental Disorders (DSM) IIIR (APA, 1987, 1994) criteria for nicotine withdrawal syndrome (Hughes and Hatsukami, 1992). These withdrawal symptoms peak dur- ing the first week of abstinence and return to baseline levels by four weeks (Hughes et al., 1990a). The intensity of these symptoms is further reduced over the course of time. The only exception to this pattern is increased weight. Weight may continue to increase over six months, and a reduction may not be seen at all or only after several months of absti- nence (Hughes et al., 1990a). A major determinant of whether nicotine is likely to be addictive is the amount and speed of nicotine delivery. The route of delivery also determines the pattern of nicotine delivery (as discussed earlier). For

256 CLEARING THE SMOKE example, each puff of a cigarette delivers a bolus dose of nicotine, result- ing in a rapid peak, which then falls to a trough level. The time between these bolus doses allows for resensitization of brain nAchRs, so that each delivery can remain reinforcing (Benowitz, 1999). In addition, this route of administration allows the delivery of a greater number or frequency of reinforcements. Other delivery routes result in a slow and persistent ab- sorption of nicotine. Subjective effects, the desire to use more of a drug, and the actual self-administration of a drug are functions of absorption rate (Henningfield and Keenan, 1993). Therefore, whereas cigarettes have high abuse potential, nicotine patches have lower abuse potential. It is also important to note that addiction to nicotine is not just a biological phenomenon, but also one in which learning or conditioning has taken place. Nicotine self-administration comes under the control of stimuli that have been associated with smoking or tobacco use. These stimuli can precipitate a strong desire for nicotine, withdrawal symp- toms, or drug effects. Exposure to these stimuli may lead to the same biological effect on neural substrates as observed from the direct actions of the drug (Childress et al., 1999). Furthermore, stimuli associated with tobacco use, such as the sensory aspects of smoking, can become reinforc- ing as well; that is, they become secondary reinforcers. In addition a to- bacco user develops expectancies regarding the use and effects of the substance, leading to a psychological reliance on the drug. The susceptibility to nicotine addiction is thus a result of both the biological effects of the drug and learning history. In addition, environ- mental factors (e.g., access to tobacco, restrictions on tobacco, social mod- eling) and genetic or organismic factors (e.g., rate of nicotine metabolism, psychiatric disorders, personality factors) may play a significant role. Spe- cific populations might be more vulnerable to nicotine addiction. Genetic twin studies have shown heritability estimates that range from 28 to 84%, with a mean estimate of 53% (Hughes, 1986). Genetic heritability has been associated with the onset as well as the persistence of smoking (Heath et al., 1998, 1999). Examples of what is inherited may be differences in sensi- tivity to nicotine (Pomerleau, 1995), the rate of nicotine metabolism (Tyndale et al., 1999), or other mechanisms such as genetic polymorphisms in the dopamine transporter and subtypes of dopamine receptors (Lerman et al., 1999; Shields et al., 1998). In addition, individuals with comorbid disorders tend to have a high prevalence of smoking. For example, high prevalence of smoking is found in individuals with depressive disorders, schizophrenia, and alcohol or drug abuse disorders (Breslau, 1995; Hughes et al., 1986a). The mechanisms responsible for susceptibility to smoking may differ across disorders. The nicotine-associated release of neurotransmitters is similar to those found with antidepressants and may be responsible for the association between smoking and depression and

NICOTINE PHARMACOLOGY 257 for the recurrence of depressive disorders after smoking cessation. Fur- thermore, studies have shown a genetic linkage between smoking and depression (Kendler et al., 1993), and observations have been made that depression can predate smoking or smoking predate depression (Breslau et al., 1993, 1998). For individuals with schizophrenia, the sensory gating effects of nicotine via the α7 nicotinic receptor may provide some symp- tomatic relief (Dalack et al., 1998; Freedman et al., 1997). A genetic link also seems to exist for alcohol and nicotine addiction (Hughes, 1986), along with commonality in the release of dopamine across all drugs lead- ing perhaps to increased sensitivity to the reinforcing effects of drugs or the potential for substitution. Furthermore, nicotine can be used to offset the aversive effects of drug use (Benowitz, 1999). Assessment of Addiction. Various measures and methods have been developed to measure dependence on a drug and its abuse or addiction potential (see Table 9-1). These measures and methods are important in examining harm reduction products since addiction to a drug is one of the determining factors associated with its harmful consequences. The addic- tive potential of a drug can be determined by examining the number of individuals, within the general population and among those exposed to the drug, who are regular users of the drug or are considered dependent on a drug, using specific criteria. Determination of the abuse potential of nicotine replacement agents has also relied on examining whether users of the product escalate their use over time or continue its use beyond a recommended period. However, deciphering whether this persistent TABLE 9-1 Measures of Dependence or Addiction and Abuse Liability Measures for Dependence or Severity of Dependence Daily or regular smoking (cotinine level) DSM criteria International Diagnostic Code Surgeon General’s report, 1988 Fagerström Tolerance Questionnaire Fagerström Nicotine Dependence Test Methods to Assess for Addiction or Abuse Liability Surveys Daily use or dependence among the general population Daily use or dependence among those exposed to the drug Escalation of drug use Relapse rates Laboratory models Psychoactive or subjective effects Drug discrimination Conditioned place preference Drug self-administration Withdrawal

258 CLEARING THE SMOKE medication use is a result of the desire to prevent relapse to cigarettes or an addiction to the product can be difficult. The “addictiveness” of a drug can also be determined by the extent to which relapse occurs among those individuals who have tried to stop using it. In addition, various animal and human laboratory methods have been developed to assess the abuse liability of a drug, including measurement of psychoactive or stimulus effects and determination of whether a drug is a reinforcer (positive or negative) leading to preference for a drug or drug self-administration (Bozarth, 1987; Balster, 1991; U.S. DHHS, 1988). According to Food and Drug Administration (FDA) guidelines, abuse liability is determined by two primary factors (see deWit and Zacny, 1995). One is the likelihood of repeated use, which is determined by the drug’s psychoactive, positive reinforcing effects and the extent to which it can relieve withdrawal symptoms as a result of chronic use. Repeated use may also be determined by the degree of unpleasant effects associated with drug use. The second factor is the incidence of adverse short- and long-term consequences as a result of use. Drugs with a greater number of adverse consequences are thought to be more likely to have abuse liability than those with fewer adverse effects. Measures and Surveys of Dependence. Surveys and instruments have been used to assess the amount and frequency of use (e.g., daily use, regular use) and whether an individual is dependent on a drug based on specific diagnostic criteria. These measurement tools have been used to determine the extent to which dependence occurs within a general popu- lation and among those who have been exposed to or have experimented with the drug. In addition, these diagnostic tools for dependence have been used to determine whether dependence on nicotine is a dose-related phenomenon. Both DSM-IIIR and DSMIV (APA, 1987,1994) and the World Health Organization (WHO) International Diagnostic Code-10 (IDC-10) (WHO, 1991) are the commonly used criteria to assess for nicotine depen- dence. According to the DSM and the IDC-10, substance dependence, including nicotine, results in several behavioral and cognitive characteris- tics and physiological manifestations (see Table 9-2). The primary criteria for dependence based on these definitions include a strong desire to take the drug for periods longer than intended, problems controlling its use, use despite negative consequences or having a higher priority than other activities or obligations, tolerance, and physical withdrawal (APA, 1994; WHO, 1991). Not all criteria have to be met, nor is any one criteria critical to satisfy a diagnosis of dependence. In the 1988 Surgeon General’s report The Health Consequences of Smoking: Nicotine Addiction, the primary criteria for drug dependence included (1) highly controlled or compulsive use of a drug, (2) psychoactive effects from the drug, and (3) drug-reinforced behavior. Additional criteria, similar to those listed in DSM-IV and IDC,

NICOTINE PHARMACOLOGY 259 TABLE 9-2 Criteria for Substance Dependence from DSM IV DSM-IV IDC-10 A maladaptive pattern of substance use, leading to clinically significant impairment or distress, as manifest by three (or more) of the following, occurring at any time in the same 12-month period Tolerance—need increased amounts of substance to Increased tolerance achieve desired effect, or diminished effect with continued use of same amount Withdrawal Sometimes, physical withdrawal Substance often taken in larger amounts or over a A strong desire to take the longer period than intended drug Persistent or unsuccessful efforts to cut down or Difficulty controlling use control substance use Great deal of time spent in activities necessary to obtain the substance or recover from its effects Important social, occupational, or recreational Higher priority given to activities given up or reduced because of substance drug use than to other use activities and obligations Substance use continued despite knowledge of having Persisting use despite a persistent or recurrent physical or psychological harmful consequences problem likely to have been caused or exacerbated by the substance SOURCE: Adapted from RCP, 2000. were also included. The number or type of symptoms experienced varies across different drugs of abuse. The major difference between nicotine and some other drugs of abuse is the lack of intoxication in regular to- bacco users that results in behavioral and cognitive disruption (U.S. DHHS, 1988). However, this makes nicotine no less an agent of addiction or dependence than other drugs (Stolerman and Jarvis, 1995). In fact, many cigarette smokers exhibit at least as many indicators of dependence as other drug users and abusers (CDC, 1995b; U.S. DHHS, 1988). Assess- ment of nicotine dependence using these criteria can be made by a number of diagnostic structured instruments including the Composite International Diagnostic Interview-Substance Abuse Module, the National Institute of

260 CLEARING THE SMOKE Mental Health-Diagnostic Interview Schedule (NIMH-DIS), and the NIMH- DIS for children (see Colby et al., 2000, for review). Other methods have been used to assess addiction or dependence on nicotine or tobacco products. For example, population surveys such as the National Household Survey on Drug Abuse (NHSDA) assess for symptoms of tobacco dependence and include such items as how many current tobacco users (1) reported daily use of the product, (2) have tried to cut down, (3) were unable to cut down or quit or experienced difficulty quitting, (4) felt a need for more tobacco for the same effect, (5) felt depen- dent, or (6) felt sick or experienced withdrawal symptoms when stopping smoking and met at least one or more of these indicators (CDC, 1995a, b; CDC, 1994). Researchers have used meeting a specified number of these symptoms as proxy measures for the DSM-IV criteria for substance de- pendence. In some assessments, individual items, such as experiencing withdrawal symptoms or difficulty quitting have been of particular focus as indicators of dependence (CDC, 1994, 1995a, b). Other reports assessing nicotine dependence determine the number of smokers who meet criteria for high level nicotine dependence accord- ing to the Fagerström Tolerance Questionnaire (FTQ; Fagerström, 1978; Fagerström and Schneider, 1989) or the revised version, the Fagerström Test for Nicotine Dependence (FTND) (Heatherton et al., 1991). Several adolescent versions have also been developed (Prokhorov et al., 1996, 1998; Rojas et al., 1998). Although this scale is continuous, a cut-off score of 6-7 or higher has been used to separate low and high level of depen- dence. Based on the measures of dependence described above, the percent- age of cigarette users that report dependence on their tobacco product varies according to the population examined (e.g., total populations, daily smokers, ever smokers, and so forth) and the definition of dependence used. According to the NHSDA, a population survey of noninstitutional- ized civilians 12 years and older, the proportion of respondents who re- ported experiencing at least one indicator of dependence was 75.2% among those individuals who used cigarettes one or more times during the 30 days preceding the survey and 90.9% among daily users (reporting daily use for ≥ 2 consecutive weeks during the 12 months preceding the survey) (CDC, 1995b). In another study, the estimated prevalence of de- pendence according to the DSM-IIIR criteria (APA, 1987) among Ameri- cans 15-54 years old sampled for the National Comorbidity Survey was about 24.1% (Anthony et al., 1994). The lifetime prevalence of dependence among middle-aged male ever smokers in Japan was 42, 26, and 32% according to IDC-10, DSM-IIR, and DSM-IV criteria, respectively (Kawakami et al., 1998). In another study, very high rates were observed with 90% of a general sample of middle-aged male smokers meeting

NICOTINE PHARMACOLOGY 261 DSM-III criteria for dependence (Hughes et al., 1987). Kandel and associ- ates (1997) used the indicators listed in the NHSDA (see above) including items assessing for frequency and quantity of use and problems related to use in order to diagnose nicotine dependence. The criteria for diagnosis were based on the DSM-IV method in which smokers must experience three or more of seven indicators of dependence. The findings showed that while 8.6% of the general population 12 years and older met criteria for nicotine dependence, 28% of those who had used tobacco products in the past year experienced nicotine dependence. A few studies have also been conducted with adolescents. The study conducted by Kandel and associates (1997) using the NHSDA examined the prevalence of nicotine dependence by age. They observed that about 19.9% of adolescents who smoked any cigarette met criteria for nicotine dependence, compared to rates ranging from 26.4 to 32.7% among smokers between the ages of 18 and 49 years and 23.7% among those 50 and older. In a study conducted in New Zealand, about 20% of a general sample of 18-year-olds were dependent on tobacco and more than half (56.4%) of the sample who smoked daily met DSM-IIIR criteria for nicotine dependence. In another survey that used the FTQ with a score of 7 or more (indica- tive of a high level of dependence, not dependence per se), only 19% of Japanese male ever-smokers age 35 and older met this criteria (Kawakami et al., 1998), but 36% of U.S. males did (Hughes et al., 1987). Among adolescent smokers, the prevalence of high level of dependence according to the FTND or FTQ has also been wide-ranging. Many of the studies assessed prevalence of high level of dependence in special populations of adolescents. The highest percentage of adolescents with a score of 7 or more on the FTQ was observed among a heavy-smoking group who par- ticipated in a nicotine patch trial, with an observed rate of 68% (Smith et al., 1996). The lowest rate was 20% using a modified FTQ with a cutoff score of 7 or higher, which was observed in vocational technical high school student smokers (Prokhorov et al., 1996). This proportion was lower than the 50% rate that the investigators observed among adult smokers. An indicator of the addiction potential of a drug is the development of daily or regular use or dependence among those who have been ex- posed to it. There is strong evidence to show that a significant number who are exposed to cigarettes may become daily smokers or dependent on them. Among high school students participating in the 1997 Youth Risk Behavior Survey (YRBS), of the 70.2% who tried cigarette smoking, 35.8% went on to smoke daily. This rate of escalation from trying cigarette smoking to regular use of tobacco is similar to the 33-50% observed in other studies (U.S. DHHS, 1994). The development of dependence among those who tried tobacco products is similarly high. In one population-

262 CLEARING THE SMOKE based study of adult smokers, about 31.9% of those who tried tobacco became dependent on it based on the DSM-III criteria (Anthony et al., 1994). In another study of young adults aged 21-30, of the 74% who had smoked tobacco at least once, 27% developed DSM-IIIR criteria for to- bacco dependence (Breslau et al., 1993). Similar data are not available for smokeless tobacco users. Existing data are limited to the number of indi- viduals who report having used smokeless tobacco in the past month versus the number who report lifetime use of smokeless tobacco; this method of calculation represents about 18% for smokeless tobacco users. This figure is compared to 37% for cigarette smokers using a similar method of calculation (U.S. DHHS, 2000). Relapse rates among those who tried to quit have been considered another indicator of dependence on or addiction to a drug. Relapse is high among a general population of smokers who have tried to quit smok- ing, with only 2.5% being able to sustain abstinence for a year (CDC, 1994). One study showed that among self-quitters, about two-thirds re- ported smoking within two days postquit (Hughes et al., 1992). The rate of relapse among a population of smokers who have undergone clinical treatment tends to be about 75%, with a significant number of these re- lapses occurring within the first few weeks. These rates and patterns of relapse are similar to those observed with smokeless tobacco (Hatsukami and Severson, 1999) and other drugs of abuse (Hunt and Matarazzo, 1973; Maddux and Desmond, 1986; Wallace, 1989). High rates of relapse are also observed among youth that smoke. Based on results from the YRBS, among high school students who smoked daily, 72.9% had ever tried to quit smoking and only 13.5% were former smokers (CDC, 1998). Most research on the dependence on nicotine replacement products has examined the persistence of use or escalation of use over time. No data are available on the prevalence of daily use in the general population or on dependence on these products according to diagnostic criteria for dependence or FTND scores. The rate of persistent use of nicotine re- placement products among smokers enrolled in clinical trials who were assigned these products is much lower than the rate of persistent use of cigarettes, ranging from 9% for nicotine gum to 18% for nicotine nasal spray (Hughes, 1998). With nicotine nasal spray the rates of persistent use are higher, and there is evidence to show that a small number escalate the amount of use over time (deWit and Zacny, 1995). In general, addiction to these products is significantly less than addiction to cigarettes due to the relatively slow absorption of nicotine, the side effects that sometimes re- sults from use, and the cost per unit of purchase. In summary, research shows that nicotine delivered via cigarettes and smokeless tobacco is likely to lead to a high prevalence of use and dependence. One third to one-half of individuals who experiment with

NICOTINE PHARMACOLOGY 263 cigarette products are likely to become regular users and dependent on them. No data are available on the initiation of nicotine replacement prod- uct use among tobacco-naïve individuals or rates of diagnosable depen- dence, although these rates are likely to be low (Shiffman et al., 1998). The number of new NRT users among those attempting to quit was approxi- mately 10% per year prior to over-the-counter (OTC) nicotine replace- ment products and 26% per year after OTC availability (Shiffman et al., 1998). Therefore, increased availability has led to increased use of these products among smokers, however, the rate of use still remains quite low. Furthermore, among smokers who use nicotine replacement products, persistent use tends to be low. Future research endeavors should concen- trate on developing uniform methods and measures for assessing nicotine dependence so comparisons can be made across products and studies. The present measures are limited to assessing the extent of dependence and limited by being designed to diagnose other drugs of abuse and not specifically to diagnose nicotine dependence. In addition, as new prod- ucts evolve, rates of initiation, regular use or persistent use and depen- dence, or progression to dependence as a result of experimentation should be assessed. Models of Addiction. Several methods have been developed using clini- cal and animal models to determine the addiction potential or abuse li- ability of a drug. These include models of self-administration, drug dis- crimination, and conditioned drug placement. Models to examine withdrawal have also been developed. For humans, subjective responses to drugs can also be determined, although these responses may not neces- sarily be associated with actual drug-taking behavior. When a drug is reinforcing, it is more likely to be self-administered or preferred compared to a control drug that has no abuse potential. The subject is exposed to a drug, typically, at varying doses and then required to choose between this particular drug and a control drug or an alterna- tive reinforcer (e.g., sucrose for animals, money for humans), or between different doses of the drug. • In self-administration models, the animal is required to perform a particular maneuver, such as lever pressing, to obtain the drug, which is typically administered intravenously. This lever pressing could be based on a fixed ratio (a specific number of responses are required prior to drug delivery), a progressive ratio (more re- sponses are required after each drug delivery), or an interval sched- ule (a certain time interval is necessary before drug delivery), or a combination of these. Scheduled reinforcement in response to en- vironmental stimuli associated with drug administration are called second-order schedules (Goldberg et al., 1981). Drugs can be made

264 CLEARING THE SMOKE available for a fixed amount of time or throughout the day. Drugs that are reinforcing prompt the subject to work more or pay a higher cost for them than for the control; reinforcing drugs also lead to a greater persistent responding for them even when they are no longer available (Henningfield et al., 1991). Typically the dose-response curve is U-shaped (Risner and Goldberg, 1983; Rose and Corrigall, 1997) with low and high doses resulting in reduced drug self-administration. Low doses may produce limited or unde- tectable effects and high doses may produce adverse effects. • Drug discrimination models involve training the subject to discrimi- nate the stimulus properties of drug A from drug B. A third drug may be introduced, and the animal or human subject is asked to decide whether the drug is more like drug A or drug B (Bigelow and Preston, 1989; Preston, 1991). Subjects can also be trained to discriminate among several sets of drugs or different doses of a drug. This model allows determination of the mechanism of action of a drug. For example, if one wanted to determine whether an opioid has µ agonist or κ agonist activity, an experiment can be developed in which the subject is trained to discriminate between drugs that are known to have each of these activities. After this period of training, the drug in question is introduced and the sub- ject has to indicate whether the drug is more like drug A (e.g., a µ agonist) or drug B (a κ agonist). This model can be also used to determine whether a drug has the stimulus properties of a particu- lar pharmacological class of drugs that are abused. A similar method is used in a drug preference procedure, in which subjects are exposed to drug A and drug B, and are required to self-administer each of these drugs during separate experimental sessions. After the drug exposure or sampling period, subjects are then asked to choose between drugs A and B, to determine their preference for one drug or another (de Wit, 1991). Drug A or B can be two differ- ent doses of a drug, different types of drugs, or an active and placebo drug. • The conditioned place preference model also is used to determine the abuse liability of a drug. Animals are trained that the drug is avail- able only in a particular place (e.g., a specific chamber). Then a determination is made of how frequently the animal is willing to go to this place. If it is chosen significantly more frequently than the other place which is associated with a control drug or no drug administration, the experimental drug may have abuse potential (Bozarth, 1987). • Drug withdrawal models have typically involved observing signs and symptoms during a period of abstinence after repeated admin-

NICOTINE PHARMACOLOGY 265 istration of a drug (U.S. DHHS, 1988; Hughes et al., 1990; Malin et al., 1992). These withdrawal symptoms can be precipitated by an- tagonist drugs or allowed to occur naturally. Although the occur- rence of withdrawal signs and symptoms does not necessarily in- dicate that that the drug will be abused, it may be one indicator of the potential for abuse. • Finally, among humans, subjective responses to drugs can be deter- mined (Jasinski and Henningfield, 1989; Fischman and Foltin, 1991; Jaffe and Jaffe, 1989). Subjects can be asked to indicate the intensity of experiencing different subjective effects, such as the degree of euphoria, liking of a drug, “high,” desire for a drug, or “head rush.” Comparisons can be made across different drugs and across doses within a particular drug. Subjects can also be asked to rate the effects of a drug using various standardized measures that have been developed to assess a drug profile (e.g., stimulant-like effects, depressant effects) such as the Addiction Research Inven- tory (Haertzen et al., 1963). Self-administration paradigms have been used to demonstrate that a wide range of species (monkeys, mice, dogs, and rats) exhibit preference for administering nicotine over a control vehicle (Henningfield and Goldberg, 1983; RCP, 2000; Rose and Corrigall, 1997; Swedberg et al., 1990; U.S. DHHS, 1988). Studies have shown that these animals are will- ing to lever-press several hundred times in order to receive an injection of nicotine (Goldberg et al., 1981; Risner and Goldberg, 1983). However, unlike other drugs such as cocaine, the range of environmental conditions under which nicotine serves as a reinforcer is more restricted (Henningfield and Goldberg, 1983). In laboratory studies, human smok- ers have also been found to lever-press for intravenous doses of nicotine (Henningfield and Goldberg, 1983) as well as to self-administer greater number of doses of nicotine nasal spray (Perkins et al., 1997) and nicotine gum (Hughes et al., 1990b) compared to the respective placebo condi- tions. Clinical trials for the nicotine spray (Sutherland et al., 1992) and gum (Hughes et al., 1991) have also observed greater self-administration of active compared to placebo doses. Most human studies, however, have focused on assessing smoking behavior, looking at various indices of exposure, including number of cigarettes, number of puffs, puff volume, puff duration, inhalation duration, and intercigarette interval as well as biochemical indices of exposure such as cotinine or nicotine concentra- tions. Smoking behavior has been examined in response to changes in dose of cigarettes, preloading with nicotine or administering nicotinic antagonists and other drugs that may affect the reinforcing effects of nico- tine (U.S. DHHS, 1988). Self-administration of nicotine is dose related in

266 CLEARING THE SMOKE both humans and animals, although there is lesser dose-dependency than other drugs in animals, and the curve is somewhat flat for humans (Corrigall, 1999). Nonetheless, reduced nicotine self-administration in humans is observed with nicotine preloading and compensation with changing nicotine doses in cigarettes. Speed of nicotine delivery also plays a role in the extent to which nicotine is self-administered. Rapid bolus injections of nicotine result in greater self-administration than a slow in- fusion (Wakasa et al., 1995). Self-administration can be blocked by mecamylamine, a nonspecific nAchR antagonist or by dopamine receptor antagonists (see earlier discussion of the biological basis of addiction). Self-administration can be facilitated not only by the dosing characteris- tics of cigarettes or nicotine but also by the sensory characteristics of cigarettes (Henningfield and Goldberg, 1983; Rose and Corrigall, 1997). Smokers tend to report dose-related subjective effects such as drug liking, drug strength, head rush, and feeling dizzy or aroused as a result of inhaled, buccal (smokeless tobacco), intravenous, or nasal spray nico- tine administration (Fant et al., 1999; Henningfield et al., 1985; Jones et al., 1999; Perkins et al., 1994a, 1994b). Smokers who have a history of drug dependence exhibit a similar dose-related increase in “liking” and other subjective responses for intravenously administered nicotine as observed for cocaine, amphetamine, morphine, pentobarbitol, and heroin (Jasinski et al., 1984; Jones et al., 1999; Keenan et al., 1994). Findings from another study also revealed that intravenous nicotine was misidentified as cocaine or amphetamine by the study participants who had histories of drug use (Henningfield et al., 1985; Jones et al., 1999). Subjective responses to nico- tine gum, patch, spray and inhaler have been less pronounced than re- sponses to cigarettes or intravenous nicotine (deWit and Zacny, 1995; Henningfield and Keenan, 1993; Schuh et al., 1997). The occurrence of withdrawal symptoms after cessation of continu- ous nicotine infusion in rodents has been demonstrated (Malin et al., 1992). In humans, withdrawal symptoms upon cigarette smoking cessa- tion has also been well established (Hughes et al., 1990a). However, fewer studies have been conducted with other tobacco products or nicotine re- placement agents. Cessation of smokeless tobacco use generally produces less intense withdrawal symptoms than cessation of cigarette smoking (Hatsukami et al., 1987; Keenan et al., 1989). However, in a population of smokeless tobacco users enrolled in clinical trials, the severity and num- ber of withdrawal symptoms from smokeless tobacco were comparable to those experienced by cigarette smokers who were trying to quit (Hatsukami et al., 2000). Nicotine gum withdrawal symptoms also tend to be significantly less intense in number and severity than cigarette with- drawal symptoms (Hatsukami et al., 1991, 1993, 1995), and higher doses of gum produce greater withdrawal than lower doses of gum (Hatsukami

NICOTINE PHARMACOLOGY 267 et al., 1991). On the other hand, among those who have used the product for a prolonged period, nicotine gum may be comparable to cigarettes in the number of withdrawal symptoms experienced (Hughes et al., 1986b; West and Russell, 1985). In summary, various laboratory studies have observed that nicotine is self-administered, produces psychoactive effects, and produces with- drawal symptoms. The route of delivery can determine the extent to which nicotine-containing products can produce these effects and lead to addic- tion, with cigarettes showing the highest potential for addiction. Future studies on new products should routinely measure the abuse potential of a drug by using the various methods that have been de- scribed. Furthermore, these paradigms could be considered to test medi- cations focused at reducing frequency of tobaccco use. Gastrointestinal Tract Nicotine exerts its effects on the gastrointestinal (GI) tract mainly via the activation of parasympathetic ganglia. Generally, it increases tone and contractility, and nausea, vomiting, and diarrhea can result from an over- dose. However, GI irritation, other than mild nausea, rarely complicates NRT (Wong et al., 1999). Salivation evoked by cigarette smoking also rarely accompanies the doses used in NRT. Nicotine slows gastric empty- ing and reduces gastric and pancreatic secretions. Given the association of smokeless tobacco with oral cancer (Schildt et al., 1998; Winn, 1997), there was initial concern that this might pose a risk with NRT. Follow-up studies of intermediate duration do not sub- stantiate this concern (Wallstrom et al, 1999). In recent years, the observa- tion that ulcerative colitis appears to be ameliorated in smokers has prompted the evaluation of NRT for this condition and controlled studies support its efficacy (Guslandi, 1999; Sandborn, 1999) and a delivery system permitting controlled release of nicotine in the colon has been developed. Other Effects of Nicotine There is much speculation about the existence of gender-specific ef- fects of nicotine and their implications for NRT strategies. There is some evidence of differences in the pharmacodynamic effects of nicotine be- tween genders and of an influence of timing in the menstrual cycle on the response to NRT and the success of attempts to quit (Pomerlau et al., 1991; Gritz et al., 1996). Women appear to have more pronounced withdrawal symptoms during the late luteal phase of the menstrual cycle, and it has been suggested that fear of weight gain, confidence in the ability to quit,

268 CLEARING THE SMOKE and readiness to quit smoking might be differentially related to gender. Maternal smoking has adverse effects on the fetus, including the risk of spontaneous abortion, abruptio placentae, reduced weight at birth, and deformities (Haustein, 1999). In animal models, maternal consumption of nicotine results in hyperactivity in the neonate (Tizabi et al., 2000). Mater- nal smoking has been associated with sudden infant death syndrome and appears to result in an intellectual deficit, apparent in children at least as old as 6-7 years of age (Frydman, 1996). Chronic smokers tend to have lower blood pressure than nonsmokers (Charlton and While, 1995). Maternal smoking has been associated with a reduced incidence of preeclampsia, but the mechanism is unclear (Lain et al., 1999). Nicotine does cross the placental barrier unchanged and mater- nal passive smoking raises nicotine levels in breast milk and in suckling infants. No linkage of nicotine consumption to birth deformities has been established; however, its contribution to the other effects of smoking dur- ing pregnancy is less clear (Haustein, 1999). For example, nicotine inhibits placental aromatase, reduces uteroplacental blood flow, and may ad- versely affect endothelial function in animal models (Torok et al., 2000). Presently, the experience with short-term NRT has not been associated with reports of adverse effects on fetal outcome, however; the number of individuals evaluated in this setting has been small. Cigarette smoking is associated with lower body weight and quitting is associated with weight gain. Involvement in a weight control program amplifies the efficacy of NRT (Danielsson et al., 1999). Although the mechanisms are likely to be complex, nicotine is of some substantial relevance to this effect of smoking. Aside from its stimulatory effect on basal metabolic rate, nicotine reduces the synthesis of neuropeptide Y (NPY) in the arcuate neurons which project into the paraventricular nucleus (PVN). Injection of NPY into the PVN results in hyperphagia and obesity in rats (Frankish et al., 1995). Smoking is associated with insulin resistance, which improves after cessation (Kong et al., 2000). Elevated leptin levels have been related to weight loss in smokers, and levels appear to correlate with the degree of insulin resistance (Assali et al., 1999). Cross- over studies in volunteers suggest that plasma leptin levels correlate with changes in insulin sensitivity and that intermediate levels are found in subjects on NRT (Oeser et al., 1999). Nicotine has diverse effects on other hormonal systems in the brain that are presently poorly understood. For example, chronic nicotine ad- ministration stimulates mediobasohypothalamic tyrosine hydroxylase and suppresses pro-opiomelanocortin mRNAs. Suppression of forebrain β-endorphins may be relevant to maintaining nicotine self-administration (Rasmussen, 1998). Similarly, smoking is extremely prevalent among schizophrenics and may modulate the response to certain antipsychotics,

NICOTINE PHARMACOLOGY 269 such as clozapine (McEvoy et al., 1999). It has been speculated that this behavioral response may represent an attempt at self-medication, and some evidence for a disease-related abnormality in central nAchR sensory gating in schizophrenia has begun to emerge (Breese et al., 2000; Dalack et al., 1998). Much less information is available concerning the effects of nicotine on other systems. Examples of potentially important observations include impairment of the immune response (Sopori et al., 1998), adverse effects on bone formation (Fung et al., 1999), and testicular hypogonadism (Kavitharaj and Vijayammal, 1999; Reddy et al., 1998), all observed with nicotine in model systems. The relevance of these observations, if any, to the doses of nicotine achieved in humans during NRT is unknown and should be evaluated. Finally, cigarette smoking may result in drug interactions. While poly- cyclic hydrocarbons in cigarette smoke induce CYP isozymes of potential relevance to carcinogenesis, nicotine itself can induce CYP2E1, CYP2A1/ 2A2, and CYP2B1/2B2 in animal studies. Cutaneous vasoconstriction due to nicotine can delay the absorption of transdermal and subcutaneously administered medication, including insulin and heparin, and the stimu- lant effects of nicotine can diminish the analgesic effects of some opioids and the sedative effects of benzodiazapines (Zevin and Benowitz, 1999). Cigarette smoking reduces the hypotensive response to β-blockers (Fox et al., 1984), but the contribution of nicotine to this effect is unknown. Smok- ing reduces portal blood flow velocity and volume in humans and may modulate the disposition of drugs subject to hepatic metabolism (Rapaccini et al., 1996). RESEARCH AGENDA NRT has proven an effective strategy in the cessation of cigarette smoking that is remarkably well tolerated at least in the short to medium term. Although the experience is much more limited, NRT also holds promise as a strategy for reducing the number of cigarettes smoked by those who cannot or will not quit. Both of these observations prompt considerations for future research. Thus, for those who quit smoking but continue to take NRT indefinitely, are there reasons to be concerned? First, nicotine can be addictive and although the daily exposure may be lower on NRT than when the indi- vidual was smoking, continued use implies psychological dependence, if not physical addiction. It is arguable whether this should be a concern, given the marked reduction of risk compared to smoking. However, it would seem reasonable to include surveillance of the dependence poten- tial and various methods to determine abuse liability of various nicotine

270 CLEARING THE SMOKE products. Furthermore, the implications of long-term nicotine intake for such factors as the safety of drug and alcohol consumption, progression of incidental diseases, impact of aging on cognitive and other physiologi- cal functions, and susceptibility to other forms of addictive behavior are largely unknown. For example, observations suggesting that nicotine im- pairs endothelial function, a property it shares with cigarette smoking, raise concerns about its effect on atherogenesis during long-term usage. Such an effect may take many years to emerge and highlights the impor- tance of continued postmarketing surveillance of NRT. This is also true of carcinogenesis. For example, nicotine may be metabolized to nitrosamines (e.g. nicotine-derived nitroketone) with carcinogenic potential (Hecht, 2001; Hoffman et al., 1991). However, the methodology to assess their formation is just emerging, and the concentration-effect relationships and individual patterns of susceptibility are far from established. Studies of long term nicotine administration on surrogate variables that more closely resemble the mechanism under consideration (e.g., imaging of plaque progression) and attendant studies in animal models seems timely. In- creasingly, the application of genomic and proteomic approaches is likely to clarify the differential effects of smoking and NRT on the expression and translation of genes related to the development of smoking-related diseases. Finally, the picture of nicotine’s effect on inflammation and the immune response is confused and limited. More research is needed to clarify its effects on cytokine generation, the formation of nitric oxide and eicosanoids and oxidative injury. Research should continue to explore other potential therapeutic efficacies of NRT, including its use in ulcera- tive colitis, analgesia, weight reduction, Parkinson’s disease, and cogni- tive disorders associated with aging and schizophrenia. Broadly speak- ing, the experience with long-term experience with nicotine via Swedish snus is reassuring with respect to safety, but formal evaluations of such risk from long-term use under controlled conditions have been scant (Idris et al., 1998; Raw and Macneil, 1990). Continued use of NRT in conjunction with ongoing, albeit reduced, smoking prompts additional questions. For example, the constituents of cigarette smoke that mediate tissue injury are largely unknown, and it is also unknown if modulating the coincident nicotine level might influence their absorption, metabolic disposition, mechanism of action, or elimina- tion. Design of such studies will rely on the development of more refined and tractable methodology to investigate the in vivo kinetics and dynam- ics of other constituents of cigarette smoke and their interactions with nicotine. Finally, although ethnicity already seems relevant, other factors that determine interindividual differences in nicotine efficacy, safety, and ad- dictive potential remain largely unexplored. Particular attention might be

NICOTINE PHARMACOLOGY 271 paid to genetic variation in proteins relevant to nicotine pharmacokinetics and dynamics and their interaction with environmental variables. As with other drugs, one anticipates increasing individualization of nicotine dos- age and/or delivery when given as a therapeutic agent. Insight into the interaction of genetic and environmental factors that influence initiation (Gynther et al., 1999; Heath et al., 1999) of cigarette smoking, latency until the practice becomes habitual (Stallings et al., 1999), and the quantity then smoked (Koopmans, 1999) has been increasing. Clarification of how these factors interact is also likely to afford insights of value in predicting the individual likelihood of response to the use of NRT as a strategy for quitting or reducing tobacco exposure. REFERENCES Adler LE, Olincy A, Waldo M, et al. 1998. Schizophrenia, sensory gating, and nicotinic receptors. Schizophr Bull 24(2):189-202. Allen SS, Hatsukami D, Gorsline J. 1994. Cholesterol changes in smoking cessation using the transdermal nicotine system. Transdermal Nicotine Study Group. Prev Med 23(2):190-196. Anthony J, Warner L, Kessler R. 1994. Comparative epidemiology of dependence on to- bacco, alcohol, controlled substances, and inhalants: basic findings from the national comobidity survey. Experimental and Clinical Psychopharmacology 2(3):244-268. APA (American Psychiatric Association). 1987. Diagnostic and Statistical Manual of Mental Disorders, Third Edition-Revised. Washington, DC: APA. APA (American Psychiatric Association). 1994. Diagnostic and Statistical Manual of Mental Disorders, Fourth Edition. Washington, DC: APA. Armitage A, Dollery C, Houseman T, Kohner E, Lewis PJ, Turner D. 1978. Absorption of nicotine from small cigars. Clin Pharmacol Ther 23(2):143-151. Assali AR, Beigel Y, Schreibman R, Shafer Z, Fainaru M. 1999. Weight gain and insulin resistance during nicotine replacement therapy. Clin Cardiol 22(5):357-360. Balster RL. 1991. Drug abuse potential evaluation in animals. Br J Addict 86(12):1549-1558. Benowitz NL. 1986. Clinical pharmacology of nicotine. Annu Rev Med 37:21-32. Benowitz NL. 1993. Nicotine replacement therapy. What has been accomplished—can we do better? Drugs 45(2):157-170. Benowitz NL. 1995. Clinical pharmacology of transdermal nicotine. Eur J Pharm Biopharm 41:168-174. Benowitz NL. 1999. Nicotine addiction. Prim Care 26(3):611-631. Benowitz NL, Fitzgerald GA, Wilson M, Zhang Q. 1993. Nicotine effects on eicosanoid formation and hemostatic function: comparison of transdermal nicotine and cigarette smoking. J Am Coll Cardiol 22(4):1159-1167. Benowitz NL, Kuyt F, Jacob P, Jones RT, Osman AL. 1983. Cotinine disposition and effects. Clin Pharmacol Ther 34(5):604-611. Benowitz NL, Gourlay SG. 1997. Cardiovascular toxicity of nicotine: implications for nico- tine replacement therapy. J Am Coll Cardiol 29(7):1422-1431. Benowitz NL, Jacob P 3rd. 1994. Metabolism of nicotine to cotinine studied by a dual stable isotope method. Clin Pharmacol Ther 56(5):483-493.

272 CLEARING THE SMOKE Benowitz NL, Jacob P 3rd. 1999. Pharmacokinetics and metabolism of nicotine and related alkaloids. Arneric SP, Brioni JD, eds. Neuronal Nicotinic Preceptors Pharmacology and Therapeutic Opportunities. Boston: Wiley-Liss, Inc. Pp. 213-234. Benowitz NL, Jacob P 3d, Jones RT, Rosenberg J. 1982. Interindividual variability in the metabolism and cardiovascular effects of nicotine in man. J Pharmacol Exp Ther 221(2):368-372. Benowitz NL, Jacob P 3d, Savanapridi C. 1987. Determinants of nicotine intake while chew- ing nicotine polacrilex gum. Clin Pharmacol Ther 41(4):467-473. Benowitz NL, Perez-Stable EJ, Fong I, Modin G, Herrera B, Jacob P 3rd. 1999. Ethnic differ- ences in N-glucuronidation of nicotine and cotinine. J Pharmacol Exp Ther 291(3):1196- 1203. Benowitz NL, Porchet H, Jacob PI. 1990. Pharmacokinetics, metabolism and pharmacody- namics of nicotine. Wonnacott S, Russell MAH, Stolerman IP, eds. Nicotine Psychophar- macology: Molecular, Cellular and Behavioural Aspects. New York: Oxford University Press. Pp. 112-157. Benowitz NL, Porchet H, Sheiner L, Jacob P 3d. 1988. Nicotine absorption and cardiovascu- lar effects with smokeless tobacco use: comparison with cigarettes and nicotine gum. Clin Pharmacol Ther 44(1):23-28. Berlin I, Said S, Spreux-Varoquaux O, Olivares R, Launay JM, Puech AJ. 1995. Monoamine oxidase A and B activities in heavy smokers. Biol Psychiatry 38(11):756-761. Bigelow GE, Preston KL. 1989. Drug discrimination: methods for drug characterization and classification. NIDA Res Monogr 92(11):101-122. Bozarth MA. 1987. Methods of Assessing the Reinforcing Properties of Abused Drugs. New York: Springer-Verlag. Brandange S, Lindbolm L. 1979. The enzyme ‘aldehyde oxidase’ is an iminium oxidase. Reaction with nicotine-∆1’(5’)-iminium ion. Biochem Biophys Res Comm 91:991-996. Breese CR, Lee MJ, Adams CE, et al. 2000. Abnormal regulation of high affinity nicotinic receptors in subjects with schizophrenia. Neuropsychopharmacology 23(4):351-364. Breslau N. 1995. Psychiatric comorbidity of smoking and nicotine dependence. Behav Genet 25(2):95-101. Breslau N, Fenn N, Peterson EL. 1993. Early smoking initiation and nicotine dependence in a cohort of young adults. Drug Alcohol Depend 33(2):129-137. Breslau N, Peterson EL, Schultz LR, Chilcoat HD, Andreski P. 1998. Major depression and stages of smoking. A longitudinal investigation. Arch Gen Psychiatry 55(2):161-166. Byrd GD, Caldwell WS, Beck DJ, Kaminski DL, Li AP. 1994 (Oct 23-27). Profile of nicotine metabolites from human hepatocytes. In: Abstract 173; ISSX Proceedings. Volume 6. Raleigh, NC: Sixth North American ISSX Meeting. Caraballo RS, Giovino GA, Pechacek TF, et al. 1998. Racial and ethnic differences in serum cotinine levels of cigarette smokers: Third National Health and Nutrition Examination Survey, 1988-1991. JAMA 280(2):135-139. CDC (Centers for Disease Control and Prevention). 1994. Reasons for tobacco use and symp- toms of nicotine withdrawal among adolescent and young adult tobacco users-United States, 1993. Morbidity and Mortality Weekly Report 43(41):745-750. CDC (Centers for Disease Control and Prevention). 1995a. Indicators of nicotine addiction among women-United States, 1991-1992. Morbidity and Mortality Weekly Report 44(6):102-105. CDC (Centers for Disease Control and Prevention). 1995b. Symptoms of substance depen- dence associated with use of cigarettes, alcohol, and illicit drugs—United States, 1991- 1992. Morbidity and Mortality Weekly Reports 44(44):830-839.

NICOTINE PHARMACOLOGY 273 CDC (Centers for Disease Control). 1998. Selected cigarette smoking initiation and quitting behaviors among high school students—United States, 1997. Morbidity and Morality Weekly Reports 47(19):386-389. Celermajer DS, Sorensen KE, Georgakopoulos D, et al. 1993. Cigarette smoking is associ- ated with dose-related and potentially reversible impairment of endothelium-depen- dent dilation in healthy young adults. Circulation 88(5 Pt 1):2149-2155. Chalon S, Moreno H Jr, Benowitz NL, Hoffman BB, Blaschke TF. 2000. Nicotine impairs endothelium-dependent dilatation in human veins in vivo. Clin Pharmacol Ther 67(4):391-397. Charlton A, While D. 1995. Blood pressure and smoking: observations on a national cohort. Arch Dis Child 73(4):294-297. Childress AR, Mozley PD, McElgin W, Fitzgerald J, Reivich M, O’Brien CP. 1999. Limbic activation during cue-induced cocaine craving. Am J Psychiatry 156(1):11-18. Chiodera P, Capretti L, Davoli C, Caiazza A, Bianconi L, Coiro V. 1990. Effect of obesity and weight loss on arginine vasopressin response to metoclopramide and nicotine from cigarette smoking. Metabolism 39(8):783-786. Chowdhury P, Hosotani R, Rayford PL. 1989. Weight loss and altered circulating GI pep- tide levels of rats exposed chronically to nicotine. Pharmacol Biochem Behav 33(3):591- 594. Cinciripini PM, Wetter DW, McClure JB. 1997. Scheduled reduced smoking: effects on smok- ing abstinence and potential mechanisms of action. Addict Behav 22(6):759-767. Colby SM, Tiffany ST, Shiffman S, Niaura RS. 2000. Measuring nicotine dependence among youth: a review of available approaches and instruments. Drug Alcohol Depend 59 Suppl 1:S23-39. Cordero-Erausquin M, Marubio LM, Klink R, Changeux JP. 2000. Nicotinic receptor func- tion: new perspectives from knockout mice. Trends Pharmacol Sci 21(6):211-217. Corrigall WA. 1999. Nicotine self-administration in animals as a dependence model. Nico- tine Tob Res 1(1):11-20. Cryer PE, Haymond MW, Santiago JV, Shah SD. 1976. Norepinephrine and epinephrine release and adrenergic mediation of smoking-associated hemodynamic and metabolic events. N Engl J Med 295(11):573-577. Dajani RM, Gorrod JW, Beckett AH. 1975. Reduction in vivo of (minus)-nicotine-1'-N-oxide by germ-free and conventional rats. Biochem Pharmacol 24(5):648-650. Dalack GW, Healy DJ, Meador-Woodruff JH. 1998. Nicotine dependence in schizophrenia: clinical phenomena and laboratory findings. Am J Psychiatry 155(11):1490-1501. Dale HH. 1914. The action of certain esters and ethers of choline and their relation to mus- carine. J Pharmacol Exp Ther 6:147-190. Danielsson T, Rossner S, Westin A. 1999. Open randomised trial of intermittent very low energy diet together with nicotine gum for stopping smoking in women who gained weight in previous attempts to quit. BMJ 319(7208):490-493; discussion 494. de Wit H. 1991. Preference procedures for testing the abuse liability of drugs in humans. Br J Addict 86(12):1579-1586. de Wit H, Zacny J. 1995. Abuse potentials of nicotine replacement therapies. CNS Drugs 4(6):456-468. Debanne SM, Rowland DY, Riedel TM, Cleves MA. 2000. Association of Alzheimer’s dis- ease and smoking: the case for sibling controls. J Am Geriatr Soc 48(7):800-806. DiChiara G. 1999. Drug addiction as dopamine-dependent associative learning disorder. Eur J Pharmacol 375(1-3):13-30. Doll R, Peto R, Boreham J, Sutherland I. 2000. Smoking and dementia in male British doc- tors: prospective study. BMJ 320(7242):1097-1102.

274 CLEARING THE SMOKE Domino EF. 1999. Pharmacological significance of nicotine. Gorrod JW, Jacob III P, eds. Analytical Determination of Nicotine and Related Compounds and Their Metabolites. Amsterdam: Elsevier. Pp. 1-11. Dugi KA, Rader DJ. 2000. Lipoproteins and the endothelium: insights from clinical re- search. Semin Thromb Hemost 26(5):513-519. Emilien G, Beyreuther K, Masters CL, Maloteaux JM. 2000. Prospects for pharmacological intervention in Alzheimer disease. Arch Neurol 57(4):454-459. Fagerström KO. 1978. Measuring degree of physical dependence to tobacco smoking with reference to individualization of treatment. Addict Behav 3(3-4):235-241. Fagerström KO, Schneider NG. 1989. Measuring nicotine dependence: a review of the Fagerström Tolerance Questionnaire. J Behav Med 12(2):159-182. Fant RV, Henningfield JE, Nelson RA, Pickworth WB. 1999. Pharmacokinetics and pharma- codynamics of moist snuff in humans. Tob Control 8(4):387-392. Fischman MW, Foltin RW. 1991. Utility of subjective-effects measurements in assessing abuse liability of drugs in humans. Br J Addict 86(12):1563-1570. Fowler JS, Volkow ND, Wang GJ, et al. 1996. Inhibition of monoamine oxidase B in the brains of smokers. Nature 379(6567):733-736. Fox K, Deanfield J, Krikler S, Ribeiro P, Wright C. 1984. The interaction of cigarette smoking and beta-adrenoceptor blockade. Br J Clin Pharmacol 17(Suppl 1):92S-93S. Frankish HM, Dryden S, Wang Q, Bing C, MacFarlane IA, Williams G. 1995. Nicotine ad- ministration reduces neuropeptide Y and neuropeptide Y mRNA concentrations in the rat hypothalamus: NPY may mediate nicotine’s effects on energy balance. Brain Res 694(1-2):139-146. Freedman R, Coon H, Myles-Worsley M, et al. 1997. Linkage of a neurophysiological deficit in schizophrenia to a chromosome 15 locus. Proc Natl Acad Sci U S A 94(2):587-592. Frydman M. 1996. The smoking addiction of pregnant women and the consequences on their offspring’s intellectual development. J Environ Pathol Toxicol Oncol 15(2-4):169-172. Fu Y, Matta SG, Gao W, Brower VG, Sharp BM. 2000. Systemic nicotine stimulates dopam- ine release in nucleus accumbens: re-evaluation of the role of N-methyl-D-aspartate receptors in the ventral tegmental area. J Pharmacol Exp Ther 294(2):458-465. Fung YK, Iwaniec U, Cullen DM, Akhter MP, Haven MC, Timmins P. 1999. Long-term effects of nicotine on bone and calciotropic hormones in adult female rats. Pharmacol Toxicol 85(4):181-187. Gairola CG, Daugherty A. 1999. Acceleration of atherosclerotic plaque formation in ApoE- mice by exposure to tobacco smoke. The Toxicologist 48:297. George TP, Verrico CD, Picciotto MR, Roth RH. 2000. Nicotinic modulation of mesoprefrontal dopamine neurons: pharmacologic and neuroanatomic characterization. J Pharmacol Exp Ther 295(1):58-66. Goldberg SR, Spealman RD, Goldberg DM. 1981. Persistent behavior at high rates main- tained by intravenous self-administration of nicotine. Science 214(4520):573-575. Gori GB, Benowitz NL, Lynch CJ. 1986. Mouth versus deep airways absorption of nicotine in cigarette smokers. Pharmacol Biochem Behav 25(6):1181-1184. Gorrod JW, Hibberd AR. 1982. The metabolism of nicotine-∆1’(5’) -iminium ion, in vivo and in vitro. Eur J Drug Metab Pharmacokinet 7:293-298. Gorrod JW, Jenner P. 1975. The metabolism of tobacco alkaloids. Hayes WJ Jr, ed. Essays in Toxicology. New York: Academic Press. Pp. 35-78. Gorrod JW, Schepers G. 1999. Biotransformation of nicotine in mammalian systems. Gorrod JW, Jacob III P, Eds. Analytical Determination of Nicotine and Related Compounds and Their Metabolites. Amsterdam: Elsevier. Gorrod JW, Wahren J. 1993. Nicotine and Related Alkaloids: Absorption, Distribution, Metabo- lism and Excretion. London: Chapman and Hall.

NICOTINE PHARMACOLOGY 275 Gouaze V, Dousset N, Dousset JC, Valdiguie P. 1998. Effect of nicotine and cotinine on the susceptibility to in vitro oxidation of LDL in healthy non smokers and smokers. Clin Chim Acta 277(1):25-37. Gourlay SG, Forbes A, Marriner T, McNeil JJ. 1999. Predictors and timing of adverse experi- ences during trandsdermal nicotine therapy. Drug Saf 20(6):545-555. Green JT, Evans BK, Rhodes J, et al. 1999. An oral formulation of nicotine for release and absorption in the colon: its development and pharmacokinetics. Br J Clin Pharmacol 48(4):485-493. Greenland S, Satterfield MH, Lanes SF. 1998. A meta-analysis to assess the incidence of adverse effects associated with the transdermal nicotine patch. Drug Saf 18(4):297-308. Gries JM, Benowitz N, Verotta D. 1996. Chronopharmacokinetics of nicotine. Clin Pharmacol Ther 60(4):385-395. Gritz ER, Nielsen IR, Brooks LA. 1996. Smoking cessation and gender: the influence of physiological, psychological, and behavioral factors. J Am Med Womens Assoc 51(1- 2):35-42. Guatura SB, Martinez JA, Santos Bueno PC, Santos ML. 2000. Increased exhalation of hy- drogen peroxide in healthy subjects following cigarette consumption. Sao Paulo Med J 118(4):93-98. Gundisch D. 2000. Nicotinic acetylcholine receptors and imaging. Curr Pharm Des 6(11):1143- 1157. Guslandi M. 1999. Long-term effects of a single course of nicotine treatment in acute ulcer- ative colitis: remission maintenance in a 12-month follow-up study. Int J Colorectal Dis 14(4-5):261-262. Guthrie SK, Zubieta JK, Ohl L, et al. 1999. Arterial/venous plasma nicotine concentrations following nicotine nasal spray. Eur J Clin Pharmacol 55(9):639-643. Gynther LM, Hewitt JK, Heath AC, Eaves LJ. 1999. Phenotypic and genetic factors in mo- tives for smoking. Behav Genet 29(5):291-302. Haertzen CA, Hill HE, Belleville, RE. 1963. Development of the Addiction Research Center Inventory (ARCI): selection of items that are sensitive to the effects of various drugs. Psychopharmacologia. 4:155-166. Hatsukami DK, Grillo M, Boyle R, et al. 2000. Treatment of spit tobacco users with transdermal nicotine system and mint snuff. J Consult Clin Psychol 68(2):241-249. Hatsukami DK, Gust SW, Keenan RM. 1987. Physiologic and subjective changes from smokeless tobacco withdrawal. Clin Pharmacol Ther 41(1):103-107. Hatsukami DK, Severson HH. 1999. Oral spit tobacco: addiction, prevention and treatment. Nicotine Tob Res 1(1):21-44. Hatsukami DK, Skoog K, Huber M, Hughes J. 1991. Signs and symptoms from nicotine gum abstinence. Psychopharmacology (Berl) 104(4):496-504. Hatsukami D, Huber M, Callies A, Skoog K. 1993. Physical dependence on nicotine gum: effect of duration of use. Psychopharmacology (Berl) 111(4):449-456. Hatsukami D, Skoog K, Allen S, Bliss R. 1995. Gender and the effects of different doses of nicotine gum on tobacco withdrawal symptoms. Experimental and Clinical Psychophar- macology 3(2):163-173. Haustein KO. 1999. Cigarette smoking, nicotine and pregnancy. Int J Clin Pharmacol Ther 37(9):417-427. Heath AC, Kirk KM, Meyer JM, Martin NG. 1999. Genetic and social determinants of initia- tion and age at onset of smoking in Australian twins. Behav Genet 29(6):395-407. Heath AC, Madden PA, Martin NG. 1998. Statistical methods in genetic research on smok- ing. Stat Methods Med Res 7(2):165-186.

276 CLEARING THE SMOKE Heatherton TF, Kozlowski LT, Frecker RC, Fagerström KO. 1991. The Fagerström Test for Nicotine Dependence: a revision of the Fagerström Tolerance Questionnaire. Br J Ad- dict 86(9):1119-1127. Hecht SS. 2001. Carcinogen biomarkers for lung or oral cancer chemoprevention trials. IARC Sci Publ. 154:245-255. Henningfield JE, Cohen C, Heishman SJ. 1991. Drug self-administration methods in abuse liability evaluation. Br J Addict 86(12):1571-1577. Henningfield JE, Goldberg SR. 1983. Nicotine as a reinforcer in human subjects and labora- tory animals. Pharmacol Biochem Behav 19(6):989-992. Henningfield JE, Heishman SJ. 1995. The addictive role of nicotine in tobacco use. Psychop- harmacology 117:11-13. Henningfield JE, Keenan RM. 1993. Nicotine delivery kinetics and abuse liability. J Consult Clin Psychol 61(5):743-750. Henningfield JE, Miyasato K, Jasinski DR. 1985. Abuse liability and pharmacodynamic char- acteristics of intravenous and inhaled nicotine. J Pharmacol Exp Ther 234(1):1-12. Hibberd AR, Gorrod, JW. 1983. Enzymology of the metabolic pathway from nicotine to cotinine, in vitro. Eur J Drug Metab Pharmacokinet. 8:151-162. Hoffmann D, Rivenson A, Chung FL, Hecht SS. 1991. Nicotine-derived N-nitrosamines (TSNA) and their relevance in tobacco carcinogenesis. Crit Rev Toxicol 21(4):305-311. Hughes JR. 1986. Genetics of smoking. A brief review. Behavioral Therapy 7:335-345. Hughes, J. 1998. Dependence on and abuse of nicotine replacement medications: an update. In Benowitz NL, ed. Nicotine Safety and Toxicity. New York: Oxford University Press. Pp. 147-157. Hughes JR, Gulliver SB, Fenwick JW, et al. 1992. Smoking cessation among self-quitters. Health Psychol 11(5):331-334. Hughes JR, Gust SW, Keenan RM, Fenwick JW. 1990b. Effect of dose on nicotine’s reinforc- ing, withdrawal-suppression and self-reported effects. J Pharmacol Exp Ther 252(3):1175- 1183. Hughes JR, Gust SW, Keenan R, Fenwick JW, Skoog K, Higgins ST. 1991. Long-term use of nicotine vs placebo gum. Arch Intern Med 151(10):1993-1998. Hughes JR, Gust SW, Pechacek TF. 1987. Prevalence of tobacco dependence and withdrawal. Am J Psychiatry 144(2):205-208. Hughes JR, Hatsukami DK. 1992. The nicotine withdrawal syndrome: a brief review and update. International Journal of Smoking Cessation 1(2):21-26. Hughes JR, Hatsukami DK, Mitchell JE, Dahlgren LA. 1986a. Prevalence of smoking among psychiatric outpatients. Am J Psychiatry 143(8):993-997. Hughes JR, Hatsukami DK, Skoog KP. 1986b. Physical dependence on nicotine in gum. A placebo substitution trial. JAMA 255(23):3277-3279. Hughes JR, Higgins ST, Hatsukami D. 1990a. Effects of tobacco abstinence. Kowzlowski LT, Annis H, Cappell, et al., eds. Research Advances in Alcohol and Drug Problems. New York: Plenum Press. Hunt WA, Matarazzo JD. 1973. Three years later: recent developments in the experimental modification of smoking behavior. J Abnorm Psychol 81(2):107-114. Idris AM, Ibrahim SO, Vasstrand EN, et al. 1998. The Swedish snus and the Sudanese toombak: are they different? Oral Oncol 34(6):558-566. Jaffe JH, Jaffe FK. 1989. Historical perspectives on the use of subjective effects measures in assessing the abuse potential of drugs. NIDA Res Monogr 92(10):43-72. Jarvik ME. 1991. Beneficial effects of nicotine. Br J Addict 86(5):571-575. Jasinski DR, Henningfield JE. 1989. Human abuse liability assessment by measurement of subjective and physiological effects. NIDA Res Monogr 92(10):73-100.

NICOTINE PHARMACOLOGY 277 Jasinski DR, Johnson RE, Henningfield JE. 1984. Abuse liability assessment in human sub- jects. Trends in Pharm Science 5:196-200. Jones HE, Garrett BE, Griffiths RR. 1999. Subjective and physiological effects of intravenous nicotine and cocaine in cigarette smoking cocaine abusers. J Pharmacol Exp Ther 288(1):188-197. Joseph AM, Norman SM, Ferry LH, et al. 1996. The safety of transdermal nicotine as an aid to smoking cessation in patients with cardiac disease. N Engl J Med 335(24):1792-1798. Kameda G, Dadmarz M, Vogel WH. 2000. Influence of various drugs on the voluntary intake of nicotine by rats. Neuropsychobiology 41(4):205-209. Kandel D, Chen K, Warner LA, Kessler RC, Grant B. 1997. Prevalence and demographic correlates of symptoms of last year dependence on alcohol, nicotine, marijuana and cocaine in the U.S. population. Drug Alcohol Depend 44(1):11-29. Kavitharaj NK, Vijayammal PL. 1999. Nicotine administration induced changes in the go- nadal functions in male rats. Pharmacology 58(1):2-7. Kawakami N, Takatsuka N, Shimizu H, Takai A. 1998. Life-time prevalence and risk factors of tobacco/nicotine dependence in male ever-smokers in Japan. Addiction 93(7):1023- 1032. Keenan RM, Hatsukami DK, Anton DJ. 1989. The effects of short-term smokeless tobacco deprivation on performance. Psychopharmacology (Berl) 98(1):126-130. Keenan RM, Jenkins AJ, Cone EJ, Henningfield JE. 1994 Nov. Smoked and intravenous nicotine, cocaine and heroin have similar abuse liability. Submitted for presentation at American Society of Addiction Medicine. Kem WR. 2000. The brain alpha7 nicotinic receptor may be an important therapeutic target for the treatment of Alzheimer’s disease: studies with DMXBA (GTS-21). Behav Brain Res 113(1-2):169-181. Kendler KS, Neale MC, MacLean CJ, Heath AC, Eaves LJ, Kessler RC. 1993. Smoking and major depression. A causal analysis. Arch Gen Psychiatry 50(1):36-43. Kong C, Elatrozy T, Anyaoku V, Robinson S, Richmond W, Elkeles RS. 2000. Insulin resis- tance, cardiovascular risk factors and ultrasonically measured early arterial disease in normotensive Type 2 diabetic subjects. Diabetes Metab Res Rev 16(6):448-453. Koob GF. 1992. Drugs of abuse: anatomy, pharmacology and function of reward pathways. Trends Pharmacol Sci 13(5):177-184. Koopmans JR, Slutske WS, Heath AC, Neale MC, Boomsma DI. 1999. The genetics of smok- ing initiation and quantity smoked in Dutch adolescent and young adult twins. Behav Genet 29(6):383-393. Labelle A, Boulay LJ. 1999. An attempted suicide using transdermal nicotine patches. Can J Psychiatry 44(2):190. Lain KY, Powers RW, Krohn MA, Ness RB, Crombleholme WR, Roberts JM. 1999. Urinary cotinine concentration confirms the reduced risk of preeclampsia with tobacco expo- sure. Am J Obstet Gynecol 181(5 Pt 1):1192-1196. Lena C, Changeux JP. 1998. Allosteric nicotinic receptors, human pathologies. J Physiol Paris 92(2):63-74. Lerman C, Caporaso NE, Audrain J, et al. 1999. Evidence suggesting the role of specific genetic factors in cigarette smoking. Health Psychol 18(1):14-20. Lopez-Arrieta JM, Rodriguez JL, Sanz F. 2000. Nicotine for Alzheimer’s disease. Cochrane Database Syst Rev (2):CD001749. Ludviksdottir D, Blondal T, Franzon M, Gudmundsson TV, Sawe U. 1999. Effects of nico- tine nasal spray on atherogenic and thrombogenic factors during smoking cessation. J Intern Med 246(1):61-66.

278 CLEARING THE SMOKE Maddux J, Desmond D. 1986. Relapse and recovery in substance abuse careers. Tims F, Leukefeld C, eds. Relapse and Recovery in Drug Abuse. Rockville, MD: U.S. Department of Health and Human Services, Public Health Service, Alcohol, Drug Abuse and Men- tal Health Administration. Malin DH, Lake JR, Newlin-Maultsby P, et al. 1992. Rodent model of nicotine abstinence syndrome. Pharmacol Biochem Behav 43(3):779-784. McEvoy JP, Freudenreich O, Wilson WH. 1999. Smoking and therapeutic response to clozapine in patients with schizophrenia. Biol Psychiatry 46(1):125-129. McKinney EF, Walton RT, Yudkin P, et al. 2000. Association between polymorphisms in dopamine metabolic enzymes and tobacco consumption in smokers. Pharmacogenetics 10(6):483-491. Meagher EA, FitzGerald GA. 2000. Indices of lipid peroxidation in vivo: strengths and limitations. Free Radic Biol Med 28(12):1745-1750. Merchant C, Tang MX, Albert S, Manly J, Stern Y, Mayeux R. 1999. The influence of smok- ing on the risk of Alzheimer’s disease [see comments]. Neurology 52(7):1408-1412. Molander L, Hansson A, Lunell E. 2001. Pharmacokinetics of nicotine in healthy elderly people. Clin Pharmacol Ther 69(1):57-65. Morrow JD, Frei B, Longmire AW, et al. 1995. Increase in circulating products of lipid peroxidation (F2-isoprostanes) in smokers. Smoking as a cause of oxidative damage. N Engl J Med 332(18):1198-1203. Murphy PJ. 1973. Enzymatic oxidation of nicotine to nicotine-∆1’(5’)-iminium ion. A newly dicovered intermediate in the metabolism of nicotine. J Biol Chem 248: 2796-2800. Nakajima M, Kwon JT, Tanaka N, et al. 2001. Relationship between interindividual differ- ences in nicotine metabolism and CYP2A6 genetic polymorphism in humans. Clin Pharmacol Ther 69(1):72-78. Nakajima M, Yamagishi S, Yamamoto H, Yamamoto T, Kuroiwa Y, Yokoi T. 2000. Deficient cotinine formation from nicotine is attributed to the whole deletion of the CYP2A6 gene in humans. Clin Pharmacol Ther 67(1):57-69. Nakajima M, Yamamoto T, Nunoya K, et al. 1996. Role of human cytochrome P4502A6 in C- oxidation of nicotine. Drug Metab Dispos 24(11):1212-1217. Niemela O, Parkkila S, Juvonen RO, Viitala K, Gelboin HV, Pasanen M. 2000. Cytochromes P450 2A6, 2E1, and 3A and production of protein-aldehyde adducts in the liver of patients with alcoholic and non-alcoholic liver diseases. J Hepatol 33(6):893-901. Nilsson P, Lundgren H, Soderstrom M, Fagerström KO, Nilsson-Ehle P. 1996. Effects of smoking cessation on insulin and cardiovascular risk factors—a controlled study of 4 months’ duration. J Intern Med 240(4):189-194. Nitenberg A, Antony I. 1999. Effects of nicotine gum on coronary vasomotor responses during sympathetic stimulation in patients with coronary artery stenosis. J Cardiovasc Pharmacol 34(5):694-699. Nomikos GG, Schilstrom B, Hildebrand BE, Panagis G, Grenhoff J, Svensson TH. 2000. Role of alpha7 nicotinic receptors in nicotine dependence and implications for psychiatric illness. Behav Brain Res 113(1-2):97-103. Nowak J, Murray JJ, Oates JA, FitzGerald GA. 1987. Biochemical evidence of a chronic abnormality in platelet and vascular function in healthy individuals who smoke ciga- rettes. Circulation 76(1):6-14. Oeser A, Goffaux J, Snead W, Carlson MG. 1999. Plasma leptin concentrations and lipid profiles during nicotine abstinence. Am J Med Sci 318(3):152-157. Oncken CA, Hardardottir H, Smeltzer JS. 1998. Human studies of nicotine replacement during pregnancy. Benowitz NL, ed. Nicotine Safety and Toxicity. New York: Oxford University Press.

NICOTINE PHARMACOLOGY 279 Oxenkrug GF. 1999. Antidepressive and antihypertensive effects of MAO-A inhibition: role of N-acetylserotonin. A review. Neurobiology (Bp) 7(2):213-224. Panagis G, Kastellakis A, Spyraki C, Nomikos G. 2000. Effects of methyllycaconitine (MLA), an alpha 7 nicotinic receptor antagonist, on nicotine- and cocaine-induced potentiation of brain stimulation reward. Psychopharmacology (Berl) 149(4):388-396. Pastrakuljic A, Schwartz R, Simone C, Derewlany LO, Knie B, Koren G. 1998. Transplacen- tal transfer and biotransformation studies of nicotine in the human placental cotyle- don perfused in vitro. Life Sci 63(26):2333-2342. Patrignani P, Panara MR, Tacconelli S, et al. 2000. Effects of vitamin E supplementation on F(2)-isoprostane and thromboxane biosynthesis in healthy cigarette smokers. Circula- tion 102(5):539-545. Perez-Stable EJ, Herrera B, Jacob P 3rd, Benowitz NL. 1998. Nicotine metabolism and intake in black and white smokers [see comments]. JAMA 280(2):152-156. Perkins KA, Grobe JE, Fonte C. 1994a. Chronic and acute tolerance to subjective, behavioral and cardiovascular effects of nicotine in humans. Journal of Pharmacology and Experi- mental Therapeutics 270:628-638. Perkins KA, Sanders M, D’Amico D, Wilson A. 1997. Nicotine discrimination and self- administration in humans as a function of smoking status. Psychopharmacology (Berl) 131(4):361-370. Perkins KA, Sexton JE, Reynolds WA, Grobe JE, Fonte C, Stiller RL. 1994b. Comparison of acute subjective and heart rate effects of nicotine intake via tobacco smoking versus nasal spray. Pharmacol Biochem Behav 47(2):295-299. Perry DC, Davila-Garcia MI, Stockmeier CA, Kellar KJ. 1999. Increased nicotinic receptors in brains from smokers: membrane binding and autoradiography studies. J Pharmacol Exp Ther 289(3):1545-1552. Pomerleau CS, Pomerleau OF, Garcia AW. 1991. Biobehavioral research on nicotine use in women. Br J Addict 86(5):527-531. Pomerleau OF. 1995. Individual differences in sensitivity to nicotine: implications for ge- netic research on nicotine dependence. Behavior Genetics. 25(2):161-177. Preston KL. 1991. Drug discrimination methods in human drug abuse liability evaluation. Br J Addict 86(12):1587-1594. Prokhorov AV, Koehly LM, Pallonen UE, Hudmon, KS. 1998. Adolescent nicotine depen- dence measured by the modified Fagerström Tolerance Questionnaire at two time points. J Child Adol Subst Abuse 7(4):35-47. Prokhorov AV, Pallonen UE, Fava JL, Ding L, Niaura R. 1996. Measuring nicotine depen- dence among high-risk adolescent smokers. Addict Behav 21(1):117-127. Quattrocki E, Baird A, Yurgelun-Todd D. 2000. Biological aspects of the link between smok- ing and depression. Harv Rev Psychiatry 8(3):99-110. Raitakari OT, Adams MR, McCredie RJ, Griffiths KA, Celermajer DS. 1999. Arterial endot- helial dysfunction related to passive smoking is potentially reversible in healthy young adults. Ann Intern Med 130(7):578-581. Raitakari OT, Adams MR, McCredie RJ, Griffiths KA, Stocker R, Celermajer DS. 2000. Oral vitamin C and endothelial function in smokers: short-term improvement, but no sus- tained beneficial effect. J Am Coll Cardiol 35(6):1616-1621. Rapaccini GL, Pompili M, Marzano MA, et al. 1996. Doppler ultrasound evaluation of acute effects of cigarette smoking on portal blood flow in man [see comments]. J Gastroenterol Hepatol 11(11):997-1000. Rasmussen DD. 1998. Effects of chronic nicotine treatment and withdrawal on hypotha- lamic proopiomelanocortin gene expression and neuroendocrine regulation. Psychoneuroendocrinology 23(3):245-259. Raw M, McNeill A. 1990. Britain bans oral snuff. BMJ 300(6717):65-66.

280 CLEARING THE SMOKE Reddy S, Londonkar R, Ravindra, Reddy S, Patil SB. 1998. Testicular changes due to graded doses of nicotine in albino mice. Indian J Physiol Pharmacol 42(2):276-280. Reilly M, Delanty N, Lawson JA, FitzGerald GA. 1996. Modulation of oxidant stress in vivo in chronic cigarette smokers. Circulation 94(1):19-25. Risner ME, Goldberg SR. 1983. A comparison of nicotine and cocaine self-administration in the dog: fixed-ratio and progressive-ratio schedules of intravenous drug infusion. J Pharmacol Exp Ther 224(2):319-326. Robinson JS, Moore VM, Owens JA, McMillen IC. 2000. Origins of fetal growth restriction. Eur J Obstet Gynecol Reprod Biol 92(1):13-19. Rojas NL, Killen JD, Haydel KF, Robinson TN. 1998. Nicotine dependence among adoles- cent smokers. Arch Pediatr Adolesc Med 152(2):151-156. Rose JE, Corrigall WA. 1997. Nicotine self-administration in animals and humans: similari- ties and differences. Psychopharmacology (Berl) 130(1):28-40. RCP (Royal College of Physicians). 2000. Nicotine Addiction in Britain. A Report of the Tobacco Advisory Group of the Royal College of Physicians. Sudbury, Suffolk: Lavenham Press, Ltd. Sabha M, Tanus-Santos JE, Toledo JC, Cittadino M, Rocha JC, Moreno H Jr. 2000. Transdermal nicotine mimics the smoking-induced endothelial dysfunction. Clin Pharmacol Ther 68(2):167-174. Sandborn WJ. 1999. Nicotine therapy for ulcerative colitis: a review of rationale, mecha- nisms, pharmacology, and clinical results. Am J Gastroenterol 94(5):1161-1171. Saxena K, Scheman A. 1985. Suicide plan by nicotine poisoning: a review of nicotine toxic- ity. Vet Hum Toxicol 27(6):495-497. Schildt EB, Eriksson M, Hardell L, Magnuson A. 1998. Oral snuff, smoking habits and alcohol consumption in relation to oral cancer in a Swedish case-control study. Int J Cancer 77(3):341-346. Schuh KJ, Schuh LM, Henningfield JE, Stitzer ML. 1997. Nicotine nasal spray and vapor inhaler: abuse liability assessment. Psychopharmacology (Berl) 130(4):352-361. Sekhon HS, Jia Y, Raab R, et al. 1999. Prenatal nicotine increases pulmonary alpha7 nicotinic receptor expression and alters fetal lung development in monkeys. J Clin Invest 103(5):637-647. Shields PG, Lerman C, Audrain J, et al. 1998. Dopamine D4 receptors and the risk of ciga- rette smoking in African-Americans and Caucasians. Cancer Epidemiol Biomarkers Prev 7(6):453-458. Shiffman S, Mason KM, Henningfield JE. 1998. Tobacco dependence treatments: review and prospectus. Annu Rev Public Health 19:335-358. Smith TA, House RF Jr, Croghan IT, et al. 1996. Nicotine patch therapy in adolescent smok- ers. Pediatrics 98(4 Pt 1):659-667. Soderpalm B, Ericson M, Olausson P, Blomqvist O, Engel JA. 2000. Nicotinic mechanisms involved in the dopamine activating and reinforcing properties of ethanol. Behav Brain Res 113(1-2):85-96. Sopori ML, Kozak W, Savage SM, et al. 1998. Effect of nicotine on the immune system: possible regulation of immune responses by central and peripheral mechanisms. Psychoneuroendocrinology 23(2):189-204. Stallings MC, Hewitt JK, Beresford T, Heath AC, Eaves LJ. 1999. A twin study of drinking and smoking onset and latencies from first use to regular use. Behav Genet 29(6):409-421. Stolerman IP, Jarvis MJ. 1995. The scientific case that nicotine is addictive. Psychopharmacol- ogy (Berl) 117(1):2-10; discussion 14-20. Strandberg TE, Tilvis RS. 2000. C-reactive protein, cardiovascular risk factors, and mortality in a prospective study in the elderly. Arterioscler Thromb Vasc Biol 20(4):1057-1060.

NICOTINE PHARMACOLOGY 281 Sutherland G, Russell MA, Stapleton J, Feyerabend C, Ferno O. 1992. Nasal nicotine spray: a rapid nicotine delivery system. Psychopharmacology (Berl) 108(4):512-518. Swahn M, Hammig B. 2000. Prevalence of youth access to alcohol, guns, illegal drugs, or cigarettes in the home and association with health-risk behaviors. Ann Epidemiol 10(7):452. Swedberg MDB, Henningfield JE, Goldberg SR. 1990. Nicotine dependency: animal studies. Wonnacott S, Russell MAH, Stolerman IP., eds. Nicotine Psychopharmacology: Molecular, Cellular and Behavioural Aspects. Oxford: Oxford Science Publications. Pp. 38-76. Tizabi Y, Russell LT, Nespor SM, Perry DC, Grunberg NE. 2000. Prenatal nicotine exposure: effects on locomotor activity and central [125I]alpha-BT binding in rats. Pharmacol Biochem Behav 66(3):495-500. Torok J, Gvozdjakova A, Kucharska J, et al. 2000. Passive smoking impairs endothelium- dependent relaxation of isolated rabbit arteries. Physiol Res 49(1):135-141. Traber MG, van der Vliet A, Reznick AZ, Cross CE. 2000. Tobacco-related diseases. Is there a role for antioxidant micronutrient supplementation? Clin Chest Med 21(1):173-187, x. Tyndale RF, Pianezza ML, Seelars EM. 1999. A common genetic defect in nicotine metabo- lism decrease smoking. Nicotine and Tobacco Research 1(S2):S61-S67. U.S. DHHS (United States Department of Health and Human Services). 1988. The Health Consequences of Smoking: Nicotine Addiction; A Report of the Surgeon General. Rockville, MD: U.S. Department of Health and Human Services, Centers for Disease Control and Prevention. U.S. DHHS (U.S. Department of Health and Human Services). 1994. Preventing Tobacco Use Among Young People; A Report of the Surgeon General. Washington, DC: U.S. Department of Health and Human Services, Centers for Disease Control and Prevention. U.S. DHHS (U.S. Department of Health and Human Services). 2000. Summary of Findings From the 1999 National Household Survey on Drug Abuse. Rockville, MD: Substance Abuse and Mental Health Services Administration. Wakasa Y, Takada K, Yanagita T. 1995. Reinforcing effect as a function of infusion speed in intravenous self-administration of nicotine in rhesus monkeys. Nihon Shinkei Seishin Yakurigaku Zasshi 15(1):53-59. Wallace BC. 1989. Psychological and environmental determinants of relapse in crack co- caine smokers. J Subst Abuse Treat 6(2):95-106. Wallstrom M, Sand L, Nilsson F, Hirsch JM. 1999. The long-term effect of nicotine on the oral mucosa. Addiction 94(3):417-423. Wang HY, Lee DH, D’Andrea MR, Peterson PA, Shank RP, Reitz AB. 2000a. beta-Amy- loid(1-42) binds to alpha7 nicotinic acetylcholine receptor with high affinity. Implica- tions for Alzheimer’s disease pathology. J Biol Chem 275(8):5626-5632. Wang H, Yang B, Zhang L, Xu D, Wang Z. 2000b. Direct block of inward rectifier potassium channels by nicotine. Toxicol Appl Pharmacol 164(1):97-101. Waters AJ, Sutton SR. 2000. Direct and indirect effects of nicotine/smoking on cognition in humans. Addict Behav 25(1):29-43. Watkins SS, Koob GF, Markou A. 2000. Neural mechanisms underlying nicotine addiction: acute positive reinforcement and withdrawal. Nicotine Tob Res 2(1):19-37. West RJ, Russell MA. 1985. Effects of withdrawal from long-term nicotine gum use. Psychol Med 15(4):891-893. Wetscher GJ, Bagchi M, Bagchi D, et al. 1995. Free radical production in nicotine treated pancreatic tissue. Free Radic Biol Med 18(5):877-882. Whiss PA, Lundahl TH, Bengtsson T, Lindahl TL, Lunell E, Larsson R. 2000. Acute effects of nicotine infusion on platelets in nicotine users with normal and impaired renal func- tion. Toxicol Appl Pharmacol 163(2):95-104.

282 CLEARING THE SMOKE WHO (World Health Organization). 1991. International classification of Disease (ICD-10). Geneva: World Helath Organization. Winn DM. 1997. Epidemiology of cancer and other systemic effects associated with the use of smokeless tobacco. Adv Dent Res 11(3):313-321. Winther K, Fornitz GG. 1999. The effect of cigarette smoking and nicotine chewing gum on platelet function and fibrinolytic activity. J Cardiovasc Risk 6(5):303-306. Wong PW, Kadakia SC, McBiles M. 1999. Acute effect of nicotine patch on gastric emptying of liquid and solid contents in healthy subjects. Dig Dis Sci 44(11):2165-2171. Yamazaki H, Inoue K, Hashimoto M, Shimada T. 1999. Roles of CYP2A6 and CYP2B6 in nicotine C-oxidation by human liver microsomes. Arch Toxicol 73(2):65-70. Yu H, Matsubayashi H, Amano T, Cai J, Sasa M. 2000. Activation by nicotine of striatal neurons receiving excitatory input from the substantia nigra via dopamine release. Brain Res 872(1-2):223-226. Zevin S, Benowitz NL. 1999. Drug interactions with tobacco smoking. An update. Clin Pharmacokinet 36(6):425-438. Zevin S, Gourlay SG, Benowitz NL. 1998. Clinical pharmacology of nicotine. Clin Dermatol 16(5):557-564.

Next: 10 Tobacco Smoke and Toxicology »
Clearing the Smoke: Assessing the Science Base for Tobacco Harm Reduction Get This Book
×
Buy Hardback | $75.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

Despite overwhelming evidence of tobacco's harmful effects and pressure from anti-smoking advocates, current surveys show that about one-quarter of all adults in the United States are smokers. This audience is the target for a wave of tobacco products and pharmaceuticals that claim to preserve tobacco pleasure while reducing its toxic effects.

Clearing the Smoke addresses the problems in evaluating whether such products actually do reduce the health risks of tobacco use. Within the context of regulating such products, the committee explores key questions:

  • Does the use of such products decrease exposure to harmful substances in tobacco?
  • Is decreased exposure associated with decreased harm to health?
  • Are there surrogate indicators of harm that could be measured quickly enough for regulation of these products?
  • What are the public health implications?

This book looks at the types of products that could reduce harm and reviews the available evidence for their impact on various forms of cancer and other major ailments. It also recommends approaches to governing these products and tracking their public health effects.

With an attitude of healthy skepticism, Clearing the Smoke will be important to health policy makers, public health officials, medical practitioners, manufacturers and marketers of "reduced-harm" tobacco products, and anyone trying to sort through product claims.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!