首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Among chronic smokers, individual differences in subjective reactions to smoking may characterize important facets of nicotine dependence that relate to abstinence-induced craving, withdrawal symptom profiles, and risk for relapse. Although the negative reinforcing properties of smoking have achieved prominent positions in models of relapse (Baker, Brandon, & Chassin, 2004), vulnerability to relapse risk may also arise from seeking positive reinforcement from smoking (Shiffman & Kirchner, 2009). In this study, 183 cessation-motivated smokers provided subjective craving, positive and negative reactions to standardized cigarettes following overnight abstinence. Level of craving, negative mood, and positive mood after overnight abstinence were significantly predictive of withdrawal on quit-day. Increased positive reactions to smoking were uniquely predictive of relapse after quitting (Hazard Ratio = 1.22, p  相似文献   

2.
Ruminative coping has been shown to heighten the risk and severity of depression. The authors hypothesized that ruminators who smoke would experience greater depressive symptoms than ruminators who do not. The rationale is that, by heightening attentional focus, nicotine may increase ruminators' ability to focus on negative thoughts, augmenting depressed mood. Participants (N = 145) self-reported smoking status, rumination, and current and lifetime depressive symptoms, including depressed mood. Results showed that rumination accounted for a larger amount of variance in current and past depressed mood and severity of lifetime depressive symptoms among smokers than nonsmokers. Noncorrelational, experimental research should directly test whether nicotine worsens depressed mood among ruminative smokers. Such evidence would be surprising because it would contradict the assumption that nicotine dispels negative moods. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

3.
Objective: Most smoking cessation studies have used long-term abstinence as their primary outcome measure. Recent research has suggested that long-term abstinence may be an insensitive index of important smoking cessation mechanisms. The goal of the current study was to examine the effects of 5 smoking cessation pharmacotherapies using Shiffman et al.'s (2006) approach of examining the effect of smoking cessation medications on 3 process markers of cessation or smoking cessation milestones: initial abstinence, lapse, and the lapse–relapse transition. Method: The current study (N = 1,504; 58.2% female and 41.8% male; 83.9% Caucasian, 13.6% African American, 2.5% other races) examined the effect of 5 smoking cessation pharmacotherapy treatments versus placebo (bupropion, nicotine lozenge, nicotine patch, bupropion + lozenge, patch + lozenge) on Shiffman et al.'s smoking cessation milestones over 8 weeks following a quit attempt. Results: Results show that all 5 medication conditions decreased rates of failure to achieve initial abstinence and most (with the exception of the nicotine lozenge) decreased lapse risk; however, only the nicotine patch and bupropion + lozenge conditions affected the lapse–relapse transition. Conclusions: These findings demonstrate that medications are effective at aiding initial abstinence and decreasing lapse risk but that they generally do not decrease relapse risk following a lapse. The analysis of cessation milestones sheds light on important impediments to long-term smoking abstinence, suggests potential mechanisms of action of smoking cessation pharmacotherapies, and identifies targets for future treatment development. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

4.
The authors examined children's depressed mood, parental depressed mood, and parental smoking in relation to children's smoking susceptibility and experimentation over 20 months in a cohort of 418 preteens (ages 10-12 at baseline) and their parents. Depressed mood in preteens was strongly related to experimentation but not to susceptibility. In cross-sectional analyses parental depressed mood was related to children's experimentation, but in longitudinal analyses parental depressed mood at baseline did not differentiate children who experimented from those who did not. Although parental smoking was strongly related to experimentation, it was not related to susceptibility either cross-sectionally or longitudinally. Depressed mood among preteens and parents appeared to be more strongly related to children's smoking behaviors than to their intentions to smoke. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

5.
[Correction Notice: An erratum for this article was reported in Vol 18(3) of Experimental and Clinical Psychopharmacology (see record 2010-11933-011). In the article the authors find it necessary to redefine the thresholding procedure used for data analyses, due to problems in the Brain Voyager software. This does not affect the main findings of the paper.] Reactivity to smoking-related cues may play a role in the maintenance of smoking behavior and may change depending on smoking status. Whether smoking cue-related functional MRI (fMRI) reactivity differs between active smoking and extended smoking abstinence states currently is unknown. We used fMRI to measure brain reactivity in response to smoking-related versus neutral images in 13 tobacco-dependent subjects before a smoking cessation attempt and again during extended smoking abstinence (52 ± 11 days) aided by nicotine replacement therapy. Prequit smoking cue induced fMRI activity patterns paralleled those reported in prior smoking cue reactivity fMRI studies. Greater fMRI activity was detected during extended smoking abstinence than during the prequit assessment subcortically in the caudate nucleus and cortically in prefrontal (BA 6, 9, 44, 46), primary somatosensory (BA 1, 2, 3), temporal (BA 22, 41, 42), parietal (BA 7, 40) anterior cingulate (BA 24, 32), and posterior cingulate (BA 31) cortex. These data suggest that during extended smoking abstinence, fMRI reactivity to smoking versus neutral stimuli persists in brain areas involved in attention, somatosensory processing, motor planning, and conditioned cue responding. In some brain regions, fMRI smoking cue reactivity is increased during extended smoking abstinence in comparison to the prequit state, which may contribute to persisting relapse vulnerability. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

6.
Stress and anxiety have been shown to increase smoking motivation. There is limited experimental data on depressed or sad mood and smoking. This study investigated the effects of two induced moods on smoking behavior. Depression scores were examined as a potential moderator and mood changes were tested as a potential mediator. Smokers (N = 121) were randomly assigned to receive either a sad induction or a neutral induction via standardized film clips. Among participants with higher depression scores, smoking duration and the number of cigarette puffs were greater in response to the sad condition. There was also a marginal interactive effect on the change in expired air carbon monoxide among this subsample; however, no differences in smoking latency or craving were observed. Changes in positive mood partially mediated the effect of condition on smoking behavior among participants with high depression scores. There was no modifying effect of gender or mediating effect of negative mood changes. The results provide preliminary support that decreases in positive mood may have a greater influence on smoking behavior among depression-prone smokers than less psychiatrically vulnerable smokers. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

7.
Aversive symptoms of abstinence from nicotine have been posited to lead to smoking relapse and research on temporal patterns of abstinence symptoms confirms this assumption. However, little is known about the association of symptom trajectories early after quitting with postcessation smoking or about the differential effects of tonic (background) versus phasic (temptation-related) symptom trajectories on smoking status. The current study examined trajectories of urge and negative mood among 300 women using the nicotine patch during the first postcessation week. Ecological momentary assessments collected randomly and during temptation episodes were analyzed using hierarchical linear modeling yielding four individual trajectory parameters: intercept (initial symptom level), linear slope (direction and rate of change), quadratic coefficient (curvature), and volatility (scatter). Early lapsers, who lapsed during the first postcessation week, exhibited more severe tonic urge and phasic negative mood immediately after quitting, and more volatile tonic and phasic urge compared to abstainers. Late lapsers, who were abstinent during the first week but lapsed by 1 month, exhibited more severe tonic urge immediately after quitting compared to abstainers. These results demonstrate the importance of early postcessation urge and negative affect and highlight the value of examining both tonic and phasic effects of abstinence from nicotine. (PsycINFO Database Record (c) 2011 APA, all rights reserved)  相似文献   

8.
Relationships among depression, alcohol use, and motivation to quit smoking were examined in a sample of 350 hospitalized smokers. Multivariate multiple regression and logistic regression analyses indicated that participants with depressed mood were more likely to have a history of problematic drinking. Participants with depressed mood and a history of problematic drinking were more likely to be nicotine dependent and anticipated greater difficulty refraining from smoking while hospitalized. Alcohol use in heavier amounts was associated with a decreased concern with negative aspects of smoking, whereas history of depression was associated with increased concern in that area. Finally, current drinking was associated with increased confidence in quitting in 1 month whereas depressed mood was associated with decreased confidence in quitting. Overall, depression and alcohol use had stronger associations with smoking-related variables than with smoking cessation motivation variables. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

9.
Acute responses to smoking are influenced by nicotine and by nonpharmacological factors such as nicotine dose expectancy and sensory effects of smoke inhalation. Because negative mood increases smoking reinforcement, the authors examined whether these effects may be altered by mood context. Smokers (n=200) participated in 2 sessions, negative or positive mood induction, and were randomized to 1 of 5 groups. Four groups comprised the 2×2 balanced placebo design, varying actual (0.6 mg vs. 0.05 mg yield) and expected nicotine dose (expected nicotine vs. denicotinized [denic]) of cigarettes. A fifth group was a no-smoking control. Smoking, versus not smoking, attenuated negative affect, as well as withdrawal and craving. Negative mood increased smoking reinforcement. However, neither actual nor expected nicotine dose had much influence on these responses; even those smokers receiving and expecting a denic cigarette reported attenuated negative affect. A follow-up comparison suggested that the sensory effects of smoke inhalation, but not the simple motor effects of smoking behavior, were responsible. Thus, sensory effects of smoke inhalation had a greater influence on relieving negative affect than actual or expected nicotine intake. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

10.
Objective: To analyze whether baseline need for cognition (NFC) was a predictor or a moderator of treatment outcome in a tailored letters intervention for smoking cessation. Design: A total of 1,499 daily smokers were recruited from general medical practices in Germany within a quasi-randomized trial testing the efficacies of two brief interventions for smoking cessation: (a) computer-generated tailored letters and (b) physician-delivered brief counseling versus assessment-only. For this study, we used data from 1,097 daily smokers who were assigned to the tailored letters or the assessment-only condition. Main Outcome Measures: self-reported 6-month prolonged abstinence from tobacco smoking assessed at 12-, 18-, and 24-month follow-ups, and smoking cessation self-efficacy assessed at 6- and 24-month follow-ups. Results: Baseline NFC predicted 6-month prolonged smoking abstinence (p = .01) and smoking cessation self-efficacy (p .05) but on smoking cessation self-efficacy (p = .05). Tailored letters resulted in higher smoking cessation self-efficacy only for persons with higher NFC. Conclusion: Higher levels of NFC are required to increase smoking cessation self-efficacy in computer-tailored interventions for smoking cessation. Considering an individual's NFC might improve the efficacy of written interventions for smoking cessation. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

11.
This study investigated predictors for smoking abstinence at 12-week follow-up among 85 smokers with a past history of alcohol dependence enrolled in a smoking cessation trial. Length of alcohol abstinence at time of enrollment and longest previous period of smoking abstinence were significantly associated with smoking status at follow-up. Multiple logistic regression with these variables entered as predictors suggested that longest previous period of smoking abstinence partially mediated the relationship between length of alcohol abstinence at enrollment and smoking status at follow-up. Additional research is warranted to identify predictors of nicotine abstinence and smoking relapse in this population and to understand the factors that mediate the relationship between length of alcohol abstinence at enrollment and smoking outcome. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

12.
Reports an error in "Brain fMRI reactivity to smoking-related images before and during extended smoking abstinence" by Amy C. Janes, Blaise deB. Frederick, Sarah Richardt, Caitlin Burbridge, Emilio Merlo-Pich, Perry F. Renshaw, A. Eden Evins, Maurizio Fava and Marc J. Kaufman (Experimental and Clinical Psychopharmacology, 2009[Dec], Vol 17[6], 365-373). In the article the authors find it necessary to redefine the thresholding procedure used for data analyses, due to problems in the Brain Voyager software. This does not affect the main findings of the paper. (The following abstract of the original article appeared in record 2009-23091-001.) Reactivity to smoking-related cues may play a role in the maintenance of smoking behavior and may change depending on smoking status. Whether smoking cue-related functional MRI (fMRI) reactivity differs between active smoking and extended smoking abstinence states currently is unknown. We used fMRI to measure brain reactivity in response to smoking-related versus neutral images in 13 tobacco-dependent subjects before a smoking cessation attempt and again during extended smoking abstinence (52 ± 11 days) aided by nicotine replacement therapy. Prequit smoking cue induced fMRI activity patterns paralleled those reported in prior smoking cue reactivity fMRI studies. Greater fMRI activity was detected during extended smoking abstinence than during the prequit assessment subcortically in the caudate nucleus and cortically in prefrontal (BA 6, 9, 44, 46), primary somatosensory (BA 1, 2, 3), temporal (BA 22, 41, 42), parietal (BA 7, 40) anterior cingulate (BA 24, 32), and posterior cingulate (BA 31) cortex. These data suggest that during extended smoking abstinence, fMRI reactivity to smoking versus neutral stimuli persists in brain areas involved in attention, somatosensory processing, motor planning, and conditioned cue responding. In some brain regions, fMRI smoking cue reactivity is increased during extended smoking abstinence in comparison to the prequit state, which may contribute to persisting relapse vulnerability. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

13.
14.
The authors examined the subtype structure of smokers classified in the precontemplation stage of change within the transtheoretical model. From a general practice-based sample of 1,499 daily smoking patients from Germany (participation rate 80%), they used a subgroup of 929 smokers who were classified in the precontemplation stage and applied latent class analysis, using the pros and cons of nonsmoking and smoking cessation self-efficacy as the defining variables. Cross-sectional validation of the emerging classes was based on smoking behavior and processes of change variables. For longitudinal validation, generalized estimation equation analyses were used on motivational and abstinence criteria from 6-, 12-, 18-, and 24-month follow-ups. A 4-class model best represented the data. Three subtypes (labeled progressive, immotive, and disengaged pessimistic) were similar to clusters identified in U.S. studies. The 4th (disengaged optimistic), by contrast, was reminiscent of a type that had previously only emerged in a Dutch study. Cross-sectional and longitudinal validation results confirmed the distinctiveness and predictive power of the classes. The findings highlight the importance of tailoring interventions for smoking behavior change to the needs of different subgroups of precontemplating smokers. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

15.
Objective: To determine the effect of adding biomarker feedback (expired air carbon monoxide) to standard quit advice on cognitive antecedents of behavior change and smoking cessation and to identify potential effect moderators and mediators. Design: Smokers (N = 160) were randomized to a control (quit advice plus leaflet) or an intervention condition (as control group plus carbon-monoxide level feedback). Cognitive measures were assessed immediately after the intervention and behavioral measures at 6 months' follow-up. Main Outcome Measures: Primary outcome measures were threat and efficacy appraisal, fear arousal, and intention to stop smoking. Secondary outcome measures were quit attempts within the last 6 months and 7-day point prevalence abstinence. Results: Threat appraisal was significantly enhanced in the intervention compared with the control group, t(158) = 2.29, p = .023, as was intention to stop smoking in the next month, t(151) = 2.9, p = .004. However, this effect on intention to stop smoking was short-lived. Groups did not differ in terms of quit attempts or abstinence at follow-up, but the intervention increased the likelihood of cessation in smokers with higher self-efficacy, χ2(1) = 5.82, p = .016. Conclusions: Carbon-monoxide level feedback enhances the effect of brief quit advice on cognitive antecedents of behavior change and smoking cessation rates but further research is required to confirm the longevity of this effect and its applicability to smokers with low self-efficacy. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

16.
Lapses within the first 2 weeks of a smoking cessation attempt are strongly associated with a return to regular smoking (S. L. Kenford et al., 1994). Unfortunately, little is known about how to prevent an initial lapse from progressing to a full relapse, and presently there are no validated lapse-responsive therapeutic interventions. The present study tested the efficacy and feasibility of rapid smoking plus counseling as a novel lapse-responsive intervention. Sixty-seven participants enrolled in a smoking treatment program involving brief counseling and a 9-week course of bupropion. Beginning on the quit day, participants' smoking behavior was tracked daily for 14 days. Once an early smoking lapse was identified, participants were randomly assigned to receive either 3 sessions of rapid smoking plus counseling or no intervention (usual care). Consistent with previous research, participants who smoked during the first 2 weeks of the quit attempt had significantly poorer 6-month outcomes (3% abstinent) than did those who did not smoke (64% abstinent). Compared with early abstainers, early lapsers were more nicotine dependent and reported greater cravings and lower confidence in their ability to abstain from smoking during the first 48 hours of abstinence. As expected, rapid smoking produced a variety of aversive effects, including increased nausea, dizziness, and vomiting as well as sharply decreased cravings to smoke. However, rapid smoking did not improve abstinence outcomes relative to usual care. Although rapid smoking has been shown to be an effective treatment for initial smoking cessation, in this preliminary study the authors failed to demonstrate its effectiveness as a lapse-responsive treatment. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

17.
This study examined baseline predictors associated with smoking abstinence among 205 smokers (113 men, 92 women) with a past history of alcoholism. Their mean age was 41.8 years, and 93% were Caucasian. Participants were randomly assigned to standard treatment (ST), behavioral counseling plus exercise (BEX), or behavioral counseling plus nicotine gum (BNIC). Factors multivariately associated with point-prevalence smoking abstinence at posttreatment (1 week after target quit date) were a longer duration of prior smoking abstinence and an interaction between treatment group and having an active 12-step sponsor. ST was more effective for those with an active sponsor, whereas both BEX and BNIC were more effective for those without an active sponsor. At 1-year follow-up, independent predictors of point-prevalence smoking abstinence were a lower Fagerstr6m Tolerance Questionnaire score (K. O. Fagerstrom, 1978) and fewer years of smoking. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

18.
Alcohol dependent smokers (N=118) enrolled in an intensive outpatient substance abuse treatment program were randomized to a concurrent brief or intensive smoking cessation intervention. Brief treatment consisted of a 15-min counseling session with 5 min of follow-up. Intensive intervention consisted of three 1-hr counseling sessions plus 8 weeks of nicotine patch therapy. The cigarette abstinence rate, verified by breath carbon monoxide, was significantly higher for the intensive treatment group (27.5%) versus the rate for the brief treatment group (6.6%) at 1 month after the quit date but not at 6 months, when abstinence rates fell to 9.1% for the intensive treatment group and 2.1% for the brief treatment group. Smoking treatment assignment did not significantly impact alcohol outcomes. Although intensive smoking treatment was associated with higher rates of short-term tobacco abstinence, other, perhaps more intensive, smoking interventions are needed to produce lasting smoking cessation in alcohol dependent smokers. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

19.
Transdermal nicotine (TN) is an efficacious smoking cessation pharmacotherapy thought to work, in part, by attenuating the effects of tobacco/nicotine abstinence and the effects of concurrently smoked cigarettes. Clinical trials suggest that TN may be less efficacious for women. This study explored the possibility of TN-related gender differences in ≥8 hour abstinent smokers (54 women, 70 men) who completed four within-subject, double-blind, placebo-controlled sessions corresponding to 0, 7, 14, and 21 mg TN. In each approximately 6.5-hr long session participants smoked an own-brand cigarette 4 hours after TN administration and physiological and subjective outcomes were examined throughout each session. Results revealed that TN suppressed some signs and symptoms of tobacco abstinence and attenuated some effects of smoking, and these effects were not dependent on gender. Women were more sensitive to the direct effects of nicotine (e.g., ratings of Nauseous) and, independent of TN dose, self-administered less nicotine when smoking and rated smoking as less rewarding. Thus, although this study does not shed light on clinical observations that TN is less effective for women, results suggest that TN might need to be combined with other interventions to supplement its effects on tobacco/nicotine abstinence and concurrent smoking. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

20.
High-trait hostility is associated with persistent cigarette smoking. To better understand mechanisms that may account for this association, we examined the effects of acute smoking abstinence and delayed versus immediate smoking reinstatement on responses to a social stressor among 48 low hostile (LH) and 48 high hostile (HH) smokers. Participants completed two laboratory sessions, one before which they had smoked ad lib and one before which they had abstained for the prior 12 hr. During each session, participants completed a stressful speaking task and then smoked immediately after the stressor or after a 15-min delay. The effect of immediate versus delayed smoking reinstatement on recovery in negative mood was significantly moderated by hostility. When reinstatement was delayed, HH participants showed significant increases in negative mood over time, whereas LH participants showed little change. When reinstatement was immediate, HH and LH smokers showed similar significant decreases in negative mood. Smoking abstinence did not moderate hostility effects. Cigarette smoking may prevent continuing increases in negative mood after social stress in HH smokers, which may partially explain their low rates of quitting. (PsycINFO Database Record (c) 2010 APA, all rights reserved)  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号