32
Electronic copy available at: http://ssrn.com/abstract=2181686 1 Social Learning Theory, Cigarette Taxes and Adolescent Smoking Behavior Mir M Ali* Abstract In this paper we evaluate the effectiveness of cigarette taxes as a mechanism to reduce smoking rates among adolescents. Using data from a nationally representative sample of adolescents, we use an endogenity-corrected model with school-level fixed effects to obtain our estimates. Moving beyond the conventional definition of smoking to a definition that recognizes the complex nature of addiction by categorizing smoking into various stages and also by controlling for peer and family effects together, we learn that adolescents are not necessarily the most responsive to taxes. Influence from peers and family plays a more significant role in influencing adolescent smoking. Forthcoming, Journal of Economics and Statistics. Keywords: Adolescent Smoking; Stages of Smoking; Social Learning Theory. Office of Regulations, Policy & Social Science, Food & Drug Administration, College Park, MD 20740, USA, phone: (240) 402-1784, fax: (301) 436-2637, e-mail: [email protected].

Social Learning Theory, Cigarette Taxes and Adolescent Smoking Behavior

Embed Size (px)

DESCRIPTION

In this paper we evaluate the effectiveness of cigarette taxes as a mechanism to reducesmoking rates among adolescents. Using data from a nationally representative sample ofadolescents, we use an endogenity-corrected model with school-level fixed effects toobtain our estimates. Moving beyond the conventional definition of smoking to adefinition that recognizes the complex nature of addiction by categorizing smoking intovarious stages and also by controlling for peer and family effects together, we learn thatadolescents are not necessarily the most responsive to taxes. Influence from peers andfamily plays a more significant role in influencing adolescent smoking.

Citation preview

  • Electronic copy available at: http://ssrn.com/abstract=2181686

    1

    Social Learning Theory, Cigarette Taxes and Adolescent Smoking Behavior

    Mir M Ali*

    Abstract In this paper we evaluate the effectiveness of cigarette taxes as a mechanism to reduce smoking rates among adolescents. Using data from a nationally representative sample of adolescents, we use an endogenity-corrected model with school-level fixed effects to obtain our estimates. Moving beyond the conventional definition of smoking to a definition that recognizes the complex nature of addiction by categorizing smoking into various stages and also by controlling for peer and family effects together, we learn that adolescents are not necessarily the most responsive to taxes. Influence from peers and family plays a more significant role in influencing adolescent smoking. Forthcoming, Journal of Economics and Statistics. Keywords: Adolescent Smoking; Stages of Smoking; Social Learning Theory. Office of Regulations, Policy & Social Science, Food & Drug Administration, College Park, MD 20740, USA, phone: (240) 402-1784, fax: (301) 436-2637, e-mail: [email protected].

  • Electronic copy available at: http://ssrn.com/abstract=2181686

    2

    1. Introduction Youth smoking is an important public health policy concern today. According to

    the latest Youth and Tobacco Use report released by the Center for Disease Control and

    Prevention (CDC) , 4000 youths aged 12-17 try their first cigarette every day in the

    United States. About 80% of adults who are smokers started smoking before they were

    21 (Liang et al., 2001) and epidemiological evidence suggests that individuals who avoid

    smoking in adolescence or young adulthood have significantly lower probabilities of

    being addicted to smoking. Even though smoking prevalence has decreased steadily over

    the years, this decrease is primarily attributable to increased quit rates rather than due to

    fewer people starting to smoke (National Center for Health Statistics, 1995). Thus,

    reducing smoking rates among adolescents remains a prime public health policy concern.

    One of the major instruments that have emerged as the main policy tool in the

    effort to reduce youth smoking is higher cigarette prices through increased cigarette

    excise taxes. Studies advocating cigarette prices to curtail adolescent smoking found that

    higher prices lead to lower levels of cigarette consumption (Chaloupka and Warner,

    2000). Regarding adolescent smoking and excise taxes, a key policy concern is the

    efficacy of tax policies aimed at changing the smoking behavior of adolescents who are at

    an initial stage of smoking or have already made a transition into addiction. In other

    words, are we only reducing smoking among adolescents for whom addiction is a less

    likely outcome, suggesting that polices to increase prices are not target efficient?

    Although an extensive literature has been devoted towards understanding the

    addictive behavior of adolescents, the question regarding the target efficiency of taxes

    has received relatively little attention. In most studies smoking has been defined as a

  • 3

    binary variable where smokers are defined as those individuals who smoked at least one

    day out of the past thirty days. This definition fails to accurately account for the different

    stages of progression a person goes through before becoming an addict, thus modeling

    the different stages of addiction only as a choice rather than an outcome. If the youth at

    the margin of deciding to take up smoking who respond to the tax policy are those with a

    lower propensity for addiction and long-term use, policy gains will be exaggerated. In

    this paper the effectiveness of price as a smoking reduction mechanism among

    adolescents is examined by stage conceptualization of smoking, i.e. instead of relying on

    the widely utilized binary indicator of smoking, we take into account the fact that an

    individuals smoking behavior evolves through a sequence of several development stages

    characterized by different smoking frequencies and intensities (Lloyd-Richardson et al.,

    2002; Mayhew et al., 2000).

    The influence of peers and parents on adolescent smoking behavior has been

    discussed in the recent literature (Christakis and Fowler, 2008; Norton et al., 1998;

    Powell et al., 2005; Powell and Chaloupka, 2005). Even though the studies found peers

    and parents play a significant role in influencing smoking among adolescent, prices still

    emerged as the main policy instrument. This could be due to the fact that most of the

    studies to date do not control for the peer and family effects simultaneously. According to

    the Social Learning Theory, adolescents acquisition behavior and values are based in

    large part on a complex web of interpersonal social relations (Lloyd-Richardson et al.,

    2002). In other words, an adolescent is influenced by his/her peer and family and such

    influences are reflected by his/her engagement in certain activities. Thus, controlling for

    both the peer and family effects (in terms of peer smoking status, parents smoking status,

  • 4

    parent-child relationship, living with both parents, etc.) together in addiction models will

    enable us to more precisely estimate the influence of price (tax) on smoking, i.e. are we

    overestimating the true impact of the price by failing to control for the peer and family

    influence?

    Despite the consistency of results across studies that have provided evidence in

    terms of the significance of peer networks in smoking decisions, the magnitude of its

    effect is still debated and questioned (Christakis and Fowler, 2008; Valente et al., 2005).

    Skepticism regarding the influence of social network has been largely due to the fact that

    much of the research was unable to address the issue of peer selection (Alexander et al.,

    2001; Hall and Valente, 2007; Hoffman et al., 2007) and thus was unable to accurately

    identify the endogenous or true social effect, i.e. the effect where an individuals

    propensity to behave in some way varies with behavior of the reference group (Manski,

    1993). Accurate estimation of this effect is relevant for policy because it suggests that

    composition of peer networks is an important determinant of adolescent smoking and that

    interventions that decline smoking propensities of individuals will also decrease the

    smoking propensities of their peers (Fletcher, 2010). Building on this existing literature

    we are able to extend our analysis in a couple of ways. First, our peer measures are drawn

    not only from the nomination of close friends, but also from the individuals who attended

    the same school as the respondents and were in the same grade with them as well, i.e.

    classmates. This will allow us to identify the differences in effects that could be exerted

    by different composition of the reference groups. Fletcher (2010) only examined the

    influence of classmates but not close friends and also did not examine the impact of

    excise taxes. Second, we are also able to build on this literature by improving upon the

  • 5

    methodology to purge the potential biases from the peer estimates, by correcting for

    endogenity while accounting for various stages of smoking addiction. Even though

    Fletcher (2010) was able to address the endogenity concerns of classmates smoking, it

    utilized the standard dichotomous measure of smoking which ignores the various stages

    of addiction that an adolescent goes through before becoming an addict.

    Modeling addiction as a stage specific phenomenon and incorporating the social

    learning theory will help us in identifying a number of important concerns that would

    need to be addressed in devising policies aimed towards reducing long-term smoking

    rates. From our analysis we would be able to identify the factors that play a more

    significant role in adolescent smoking behavior. We would also be able to identify the

    role that prices, state policies, peer effects and family influence play at different stages of

    addiction. Although it might be correct to assume that it is irrelevant whether higher taxes

    directly result in less tobacco consumption or whether the effect is transmitted through

    the behavior of peers - this assumption would likely hold true for adult smokers only.

    Most states in the U.S. prohibits sale of cigarettes to individual under the age of 18 (i.e.

    adolescents), thus it is important from a policy perspective to identify the main

    mechanism of influence so that other policies can be devised to complement excise taxes.

    For example, if peer influence is identified to be a major contributor to smoking then it is

    quite likely that an adolescent experimented with tobacco not because of the price of the

    cigarette but because he/she might have seen their friends smoke.

    In sum, in this paper we aim to (i) analyze the effectiveness of tobacco tax

    policies as a tool to reduce smoking rates. In other words, we intend to find out if these

    price measures are an overestimate and indicate target inefficiencies, i.e. reduce smoking

  • 6

    among adolescents for whom addiction is a less likely outcome; (ii) analyze how

    different factors (price increase along with state policies, family influence, association

    with friend) affect smoking behavior at various stages of addiction, and (iii) propose a

    mechanism to obtain unbiased peer effect estimates.

    2. Empirical Model and Data

    2.1 Data

    The data utilized in our study are drawn from The National Longitudinal Study of

    Adolescent Health (Add Health). Add Health consists of data on adolescents in 132

    schools nationwide between grades 7 to 12. The in-school portion of the first wave of the

    survey (1994) contains cross-section data on about 90,000 adolescents. A subset of the

    initial sample (20,745 respondents) was also interviewed in their home (in-home portion

    of the data). The primary data for our analysis come from the first wave of the in-home

    survey portion of Add Health. Parents were also interviewed in the first wave of the in-

    home sample. This allows us to control for a wide range of parent-child relationship

    measures as well as the smoking status of the parents.

    A primary advantage of Add Health is that it asked respondents to nominate their

    five closest male and five closest female friends and since these friends were also part of

    the survey we were able to construct peer measures of smoking from the responses of the

    friends themselves. The average number of nominated friends per individual is 2.54 and

    approximately 85% of the friends are from the same school as the respondent. Thus, the

    sample of our analysis with nominated peers consists of 6,549 adolescents with at least

    one nominated friend interviewed in Add Health. The sample size of our grade-level peer

    analysis consists of 19,988 individuals. Besides smoking information, Add Health

  • 7

    contains an extensive range of variables regarding smoking related state policies and

    various demographic variables. Table 1 reports descriptive statistics from the first wave

    (1994) of Add Health data.

    2.1.1. Categorical Smoking Variables

    Our main dependent variable in the study is the individual's smoking stage. The

    smoking stages were created based on the individuals smoking frequency and recency .

    Following Llyod-Richardson et al. (2002) we categorize people into the following stages

    of smoking addiction:

    (i) Never Smoker - Those respondents who denied ever trying a puff of cigarettes. Such

    adolescents who have never smoked are either unaware of positive reasons to initiate

    smoking or are ignoring or resisting pressure to smoke (Mayhew et al., 2000).

    (ii) Experimental - Those who endorse trying cigarettes, although denied smoking within

    the past 30 days or ever smoking regularly (i.e. daily smoking). This stage is marked by

    adolescents trying their first few cigarettes.

    (iii) Intermittent - Those who reported smoking between 1 and 29 out of the past 30 days.

    This stage is characterized by a gradual increase in the frequency of smoking and an

    increase in the variety of situations in which cigarettes are used. Adolescents

    in this stage have progressed beyond sporadic smoking to smoke on a higher but still

    infrequent basis.

    (iv) Regular - Those who responded smoking on a daily basis within the past thirty days.

    In this final stage adolescent may experience nicotine dependence, withdrawal symptoms

    and may find it difficult to quit.

  • 8

    In the previous literature, the most widely used measure of smoking was a

    dichotomous variable that categorizes smokers as anyone who reported smoking at least

    one day out of the past thirty days or a continuous variable measuring the number of

    cigarettes smoked (Chaloupka and Warner, 2000). Tax estimates on such variables might

    be an overestimate since it would fail to control for the various stages of addiction that a

    person might go through. From our definition of smoking stages we can see that a person

    might have smoked zero days out of the past thirty days and yet could fall into the

    category of someone who has experimented with cigarettes and has the possibility of

    making a transition into being a regular smoker. Thus, an overestimation of price is quite

    likely without distinctions being made among the various stages of addiction since the

    experimental smokers will be categorized as non-smokers.

    2.1.2. Peer Measure Variables

    We construct two different measures of peer variables. One is based on the

    respondents nomination of close friends and another is based on the people who were in

    the same grade and school as the respondent. For the nominated friends we created a

    variable pertaining to the percentage of friends who are smokers, i.e. they are either

    experimental, intermittent or regular smokers. The school-level peer smoking measure is

    the percentage of students (excluding the respondent) in the same grade and school as the

    respondent who are smokers, i.e. they are either experimental, intermittent or regular

    smokers.

    2.1.3 Parent Measures

    Since Add Health interviewed one of the parents, we can control for the parent-

    child relationship via not only how the child perceives it to be, but also how the parent

  • 9

    perceives it. This will allow us to capture with more precision the role that parents play in

    influencing an adolescents smoking behavior (Ennett at al., 2001). Besides the smoking

    behavior of the parents, it is also the relationship that the child has with his/her parents

    that is expected to play a vital role in influencing their smoking behavior. Apart from

    controlling for the respondent's perception of whether he/she thinks that his/her parents

    care, understand, and pay attention, we also control for the respondents' parents

    perception of whether they think they get along well with their children, whether they

    feel they can trust their children, etc.

    2.1.4 Policy Measures

    Besides cigarette excise taxes (in dollars) an extensive array of state policy

    measures are controlled for. Like taxes, state policy measures have been labeled as one of

    the most effective policy instruments that could be utilized to curtail smoking among

    adolescents. State policies such as banning cigarettes sales via vending machine,

    marketing restrictions on billboards, prohibition of distribution of free samples as

    promotional tools, state initiating dissemination of information regarding the adverse

    effect of smoking etc. are all likely to contribute to preventing adolescents from taking up

    smoking. Besides these, various other state and local enforcement programs are likely to

    limit the availability of cigarettes among adolescents.

    In addition to these measures, we also control for socio-demographic factors like

    age, the grade they are in, family structure and gender.

    2.1.5 Factor Analysis

    Following DeCicca et al. (2007) and Kan and Tsai (2004), instead of relying on

    multiple indicators to measure the parent-child relationship and state policy measure, we

  • 10

    created an index to control for that. The indicators of the parent-child relationship and

    state policy were obtained after conducting exploratory (principal factor) factor analysis.

    This is a data reduction technique conducted to minimize correlated regressor problems,

    collinearity and improve interpretation of mutually exclusive components of responses.

    Thus, factor analysis was utilized as a statistical technique to find an indicator(s) of

    orthogonal common factor that will linearly summarize a set of original variables related

    to the parent-child relationship and state policies (Kan and Tsai, 2004). Even though an

    interpretation of these indexes as estasticities is not advisable, this would allow us to infer

    from the estimates whether the parent-child relationship and state policies have a

    significant position or negative effect on adolescent smoking behavior.

    We utilized the Kaiser criterion since it is one of the most utilized methods in the

    literature in deciding the number of factors to be retained. According to the Kaiser

    criterion, only factors that have eigenvalues greater than 1.0 are retained. Table 2 through

    4 reports the factor analysis results. The factor loadings and scoring coefficients of the

    factors with eigenvalues greater than 1.0 are presented in Table 3 and 4. From Table 2,

    we can see that for both the parent-child relationship and state policy variables there is

    only one factor with an eigenvalue greater than 1. This suggests that only one factor is

    sufficient to summarize all the variables pertaining to the parent-child relationship and

    anti-smoking state policies. Thus in our regressions we include this single factor as an

    index for parent-child relationship and anti-smoking state policies.

    2.2 Estimation Framework

    We estimate an ordered probit model of smoking behavior where the dependent

    variable, *iY representing an individuals smoking stage (0 = never smoker; 1 =

  • 11

    experimental smoker; 2 = intermittent smoker; 3 = regular smoker) is specified as

    follows:

    iiiiiii SXCFPY 543210* (1) where Pi refers to the vector of peer measures (nominated friends and school-based) for

    individual i, Fi refers to individuals vector of family characteristics such as parent-child

    relationship, parents smoking status and easy access to cigarettes at home, Ci refers to

    state level cigarette taxes faced by the individual and state level tobacco control policies.

    Xi refers to the vector of demographic characteristics such as age, gender, race, education

    etc. and finally Si is the school dummies that controls for unobserved school type

    (school-level fixed effects) or school-level factors that are common to all individuals

    within the same school. Finally, i refers to the error term. Given the standard normal assumption for i we simply compute each response probability as follows:

    )()|()|()|0( 111* xxxPxYPxYP

    *2 121( 1| ) ( | ) ( ) ( )P Y x P Y x x x

    *3 232( 2 | ) ( | ) ( ) ( )P Y x P Y x x x

    )(1)|()|3( 33* xxYPxYP

    where 321 are threshold parameters and x is the vector of all of the explanatory variables.

    The main coefficient of interest is the endogenous effect 1, which indicates the extent of peer influence on an individuals decision to smoke. If 1 is estimated to be positive, then any policy intervention that alters the smoking behavior of the individual

    within a reference group or social network would have an effect on non-treated

  • 12

    adolescents smoking behavior that are in the same social network (Ali et al. 2011). A

    standard linear regression using average contemporaneous measure by a reference group

    as a proxy for social interactions is easy to estimate. However, such measures of peer

    networks or social interactions have quite a few problems of interpretation (Ali et al.,

    2011). Manski (1993, 2003) identifies and distinguishes between three different types of

    peer or social effects:

    a. Endogenous This effect occurs when individual behavior responds to the behavior of

    others in the reference group.

    b. Exogenous or Contextual effect This occurs when individual behavior responds to

    the exogenous characteristics of the reference group.

    c. Correlated effect This occurs when individuals in the same group behaves similarly

    because they have similar unobserved characteristics or they face similar institutional

    characteristics. Therefore, correlated effects do not signify social interactions, but

    confound its estimation.

    In the case of smoking, endogenous effect can take place if individuals are more

    likely to smoke if their classmates or close friends smoke, i.e. their decision to participate

    in smoking is interrelated. Contextual (or exogenous) effect can occur, for example, if an

    individual is more likely to smoke if he or she is surrounded by peers who have parents

    who smoke (Ali et al., 2011). Correlated effects can occur if individuals in the same

    school or neighborhood choose to smoke because they each face a lower opportunity cost

    of smoking, for example, low cigarette prices, easy availability etc. This could also occur

    if individuals with similar taste or preference for smoking tend to be friends, i.e. peer

    selection (Alexander et al., 2001; Hall and Valente, 2007; Hoffman et al., 2007).

  • 13

    Manski (1993, 2000) also notes that while both the endogenous and exogenous

    effects signify social interactions, correlated effects, however, are a statistical non-social

    phenomenon. As such, policy implications would differ significantly depending on the

    type or category of the social effect that we are identifying. For example, if smoking in

    schools is characterized by an endogenous effect then any policies that are designed to

    reduce smoking in one particular group or an individual will produce a social multiplier

    and affect individuals who were not directly targeted by the policy (Ali et al., 2011).

    Exogenous effects on the other hand may not imply the same multiplier effect since the

    policy does not alter the reference groups characteristics. Correlated effects, which are

    not evidence of social interactions, can lead to misleading policy implications if they are

    not identified or adequately controlled for. Standard regressions of individual propensity

    to engage in a particular activity on group means are unable to distinguish between the

    endogenous, exogenous and correlated effects and thus is unable to provide a causal

    estimation of peer effects. This identification difficulty, coined as the reflection problem

    by Manski, occurs because group behavior by definition is the aggregation of individual

    behavior, i.e. group behavior affects individual behavior and vice versa due to the

    simultaneity in choices. Thus, for the purpose of devising effective policy it is important

    to purge these biases from peer effect estimates to identify whether peer influence is more

    important than peer selection (Ali et al., 2011; Norton et al., 1998).

    Estimating our models with Si, the school-level fixed effects, potentially mitigates

    the correlated effects. However, an instrumental variable regression is also necessary in

    this empirical analysis because of the reflection problem. The reflection problem, as

    mentioned previously, arises because peer behavior affects individual behavior and vice

  • 14

    versa. Manski demonstrated that most estimates of 1 are not identified without utilizing instrumental variables or other similar methodologies (Manski, 1993, 2000). The

    instrumental variable solves the reflection problem by identifying variables that are

    related to the group level variable but not to the individual level variable. To utilize the

    instrumental variable technique or IV one must first find variables or instruments that

    have two properties. First, they have to be correlated with (causes variation in) the

    variable whose effect we want to know about. Here it is the peer measure. Second, these

    instruments must have no direct effect on the outcome measure (Yi in 1). For our

    instrument we propose three variables:- (i) the percentage of peers who have parents who

    smoke, (ii) the percentage of peers who live with both biological parents and (iii) the

    percentage of peers who have easy access to cigarettes at home. These peer level

    variables do not predict individual behavior but rather predicts the peer level behavior.

    The intuition behind the instruments is that, while individuals who have parents who

    smoke are more likely to smoke, the proportion of individuals friend who have smoking

    parents will only directly affect the friend but not the individual. Similar intuition applies

    to the other instruments; they affect the individuals friends decision to smoke but not

    that of the individual himself or herself. The first stage will thus estimate the following

    equation -

    iiiiii XCFZP 43210 (2) where iZ refers to the instruments and i refers to the residual. However, we

    cannot implement the two-stage least squares estimation due to the ordered and

    categorical nature of the dependent variable (Terza et al., 2008). We thus implement a

    control function approach in the ordered probit regression to address the issue of

  • 15

    endogenity (Lee, 2007; Terza et al., 2008; Fang et al., 2009). In the first stage of the

    control function approach, a reduced form regression is estimated and the result is used to

    generate predicted value of the endogenous variable (peer smoking in our case). This

    predicted value is then used to obtain the fitted residual, which is then included as an

    additional regressor (along with the endogenous variable) in the second-stage estimation.

    Terza et al. (2008), strongly advises about the inclusion of this residual in estimation of

    nonlinear models with endogenous regressor and provides evidence that this approach

    leads to more consistent results.

    It is possible that the first stage of this control function approach could be an

    ordinary least square (Lee, 2007; Fang et al., 2009). After obtaining the estimated

    coefficients, we can predict the dependent variable in Equation (2) as P . We then obtain

    the fitted residual in Equation (2) by subtracting the predicted value P from P. The second stage inserts the fitted residual into Equation (1) and estimates the following

    model by ordered probit regression:

    iiiiiii SXCFPY ^

    6543210* (3)

    Due to the two-step feature of the model, we adjust the standard errors in the

    second step by nonparametric bootstrap techniques using 200 replications. Combined

    with the school-level fixed effects, the control function approach will enable us to obtain

    evidence suggestive of a causal effect of social interaction on individual smoking

    behavior. We also undertake several tests to verify the validity of our instruments.

    3. Empirical Results

    Table 5 presents our ordered probit model without the social learning mechanism.

    From the table we can see that even without controlling for social networks, cigarette

  • 16

    taxes do not have a significant effect on all smoking stages and affects only the never

    smokers and regular smokers. This suggests that the dichotomous measure of smoking

    status could lead to an upward bias in price estimates. Experimental smokers could

    progress on to becoming regular smokers and an increase in cigarette excise tax alone

    might not be able to prevent the progression to a higher addictive stage. The primary

    reason for adolescents being an important group for policy focus is to prevent them from

    initiating smoking and reduce eventual addiction. This may not be accomplished with

    reliance on prices alone. Also, since the effect of tax on experimental and intermittent

    smokers are statistically insignificant, it could imply that taxes alone may not be able to

    prevent the progression of such smokers into the full addictive stage.

    The policy index is also significant for the never smokers and regular smokers

    only. Thus, state level policy measure might induce a reduction in smoking by the regular

    smokers or induce a quitting behavior. It might also prevent the never smokers from

    taking up smoking. The taxes and policy index being significant and greater in magnitude

    for the never smokers confirms our prior hypothesis that tax policies in particular might

    not be target efficient, i.e. unable to reduce smoking among those with a higher

    propensity for addiction. Never smokers are the ones with the lowest propensity of

    transitioning into an addictive stage and taxes have the greatest effect on them.

    Demographic variables, parents education and work status has the expected effects.

    Older individuals and those who are in a higher grade are less likely to be never smokers;

    whites are more likely to be smokers than blacks and Hispanics. Being religious

    (frequency of religious service attendance) is negatively associated with smoking.

  • 17

    Parents education is also negatively related with smoking status, whereas having parents

    who work full-time outside the home is positively related to smoking.

    From Tables 6 and 7 we can see that the effect of taxes and state policy index

    becomes statistically insignificant after controlling for peer and family smoking (i.e.

    social learning theory). After the inclusion of peer smoking, parent smoking, parent-child

    relationship index and other variables related to the social learning mechanism, cigarette

    excise tax is no longer significant for any stage of smoking. The peer measure in Table 6

    pertains to close friends and the peer measure in Table 7 pertains to classmates. In both

    cases, peer and family factors appears to play a significant role, with the classmates peer

    measure being larger in magnitude compared with close friends peer measure. However,

    as mentioned previously in Section 2.2, the peer measures are likely to be endogenous.

    We adopt a control function approach to account for that in Tables 8 and 9. For the

    purpose of brevity we show the results here for our main variables of interest only, since

    the effect of the other control variables are quite similar to those presented in Tables 5 to

    7.Appendix 1 provides estimates of our first stage regression of the control function

    approach.

    From Table 8 we see that the peer effects are only statistically significant for

    never smokers The estimated peer coefficient means that a 10% increase in friends who

    are smokers decreases the probability of being a never smoker by 3%. The statistically

    insignificant effect of peer smokers underscores the importance of controlling for

    endogenity of peer smoking. However, parental characteristics such as parental smoking

    status, parent-child relationship, easy access to cigarettes exert a significant influence on

    adolescent smoking behavior as hypothesized. These results are in line with our

  • 18

    hypothesis that the social learning mechanism will have different effects on different type

    of smokers and that taxes might not be as effective a policy instrument as the previous

    literature has estimated it to be. Particularly, in this case, our results may imply that

    parental smoking behavior might be more influential, especially for experimental,

    intermittent and regular smokers

    It is important to mention that models that regress individual smoking stage on

    friends smoking stage without accounting for schoollevel fixed effects produce a

    larger peer effect estimates for all stages of smoking. However, such peer effect estimates

    are biased since they are unable to purge the correlated effects from the peer measures to

    identify the true (endogenous) peer effect. Estimating models with school-level fixed

    effects reveals that individuals that are smokers might be more highly influenced by

    unobserved environmental factors in their surroundings and this argument is further

    advanced by the positive and significant effect of easy access to cigarettes on all stages of

    smoking. In this regard, we might expect the more exogenous school-level peer measure

    to exert a greater influence on being a smoker. Conceptually, school-level peer measures

    are likely to operate through different causal mechanism than nominated peers. For

    example, such broader measures of peer social network are likely to operate through the

    establishment of smoking norms rather than operating through influence. Thus

    unobserved school-level factors that are common to all individuals within the same

    school could be more important to being a smoker.

    With peers defined as students in the same grade and school as the respondent,

    excluding the respondent, we see from Table 9 that the peer measures are greater in

    magnitude compared to the nominated peer measures, except for experimental smokers,

  • 19

    where peer effects are not statistically significant. This greater effect of classmates over

    nominated peers imply that school based intervention strategies to prevent smoking might

    be quite effective since a change in behavior in one adolescent might spread to all the

    classmates regardless of whether they are close friends or not. Cigarette excise tax along

    with state-level policy index, are once again not significant, implying that for adolescents

    the social learning mechanism could be more important, even though its effect could vary

    in magnitude depending on the stage of addiction. Another important thing to note from

    these sets of results is that the coefficient of our fitted residual is significant at the 1%

    level for both nominated friends and classmates, indicating the importance of accounting

    for the endogenity of peer smoking.

    4. Conclusion

    Prices or an increase in excise taxes have emerged as the main policy instrument

    to reduce adolescent smoking. The vast majority of the literature has found prices to have

    a significant negative impact on the probability of being a smoker. In this study we

    analyze the effectiveness of prices (tax) as a smoking reduction mechanism by estimating

    models that incorporate the stage conceptualization models of smoking and Social

    Learning Theory. Biases in peer social networks have been accounted for by using

    school-level fixed effects in the instrumental variable or the two-stage least square

    analysis and two alternative definitions of peer, nominated friends and school-level peers.

    Moving beyond the conventional definition of smoking to a definition that

    recognizes the complex nature of addiction by categorizing smoking into various stages,

    we learn that prices are necessarily not the most effective instrument to curtail smoking.

  • 20

    From the stage conceptualization of smoking model specification we learn that the effect

    of prices across the different stages is not homogeneous even when we do not control for

    the social learning mechanism. Prices remain insignificant after controlling for the social

    learning mechanism under both peer measures. A possible avenue for future research

    could be to evaluate how excise tax influence adult smokers using the stage

    conceptualization of smoking rather than a binary indicator to measure smoking behavior.

    A double-hurdle model (Jones 1989) could be adopted to examine this issue since it

    would allow for the separation of non-starters and quitters. As with any empirical

    strategy, our approach is subject to criticism and thus it is prudent to regard our results as

    demonstrating a strong association between peer smoking and individuals smoking

    behavior at various stages of addiction rather than demonstrating a causal relationship. If

    more evidence from future studies confirms our findings, the resulting body of literature

    may lead readers to infer causality. Our study only points in that direction provided that

    our assumptions hold.

    Adolescent smoking behaviors are not only dependent on their own propensity to

    be an addict, but are also dependent on an extensive array of learning factors that are

    influenced by their peer and parents. The subjective judgment on the cost and

    benefits of smoking plays a very significant role in smoking decision. An individuals

    prior belief on risk and benefits from smoking is a crucial factor in the smoking decision,

    which depends largely on his demographic and other socio-economic characteristics

    (Lahiri and Song, 2000). The Social Learning Theory states that the interpersonal social

    relationship in which the individual is fixed is of primary importance in adolescent

    acquisitions of behaviors and values.

  • 21

    From our results we can see that relying on taxes alone is not sufficient to reduce

    smoking. Recent findings by Katzman et al. (2007) on the existence of a social market

    through which adolescents can obtain cigarettes further reinforce our conclusion.

    Katzman et al. (2007) also concludes that prices will be ineffective in influencing

    smokers who obtain cigarettes through the social market. Increases in prices will not

    achieve target efficient reductions of adolescent smoking among those who continue to

    be around peers who smoke and live in households where parents smokes and access to

    cigarettes is easy. Thus, increased social awareness regarding the adverse effect of

    smoking is more likely to contribute towards a reduction in smoking.

    Acknowledgement

    The author would like to thank Debra Dwyer (Stony Brook University), John Rizzo (Stony Brook University) and Mark Montogomery (Stony Brook University) for their helpful comments and suggestions. The views expressed in this paper are those of the author and do not necessarily reflect the view of the Food & Drug Administration. This research uses data from Add Health, a program project designed by J. Richard Udry Peter S. Bearman, and Kathleen Mullan Harris, and funded by a grant P01- HD31921 from the National Institute of Child Health and Human Development, with cooperative funding from 17 other agencies. Special acknowledgment is due Ronald R. Rindfuss and Barbara Entwisle for assistance in the original design. Persons interested in obtaining data files from Add Health should contact Add Health, Carolina Population Center, 123 W. Franklin Street, Chapel Hill, NC 27516-2524 ([email protected]).

  • 22

    References

    Alexander, C., Piazza, M., Mekos, D., and Valente, T. (2001), Peers, Schools, and Adolescents Cigarette Smoking. Journal of Adolescent Health, 29(1): 22 30. Ali. MM., Amialchuk, A., and Dwyer, DS. (2011), The Social Contagion Effect of Marijuana Use among Adolescents. PLoS ONE 6(1): e16183. Chaloupka, F., and Warner, K. (2000), The Economics of Smoking. In Newhouse J. Culyer, A. (eds), Handbook of Health Economics Volume 1B. North- Holland, 2000. Christakis, N., and Fowler J. (2008), The Collective Dynamic of Smoking in a Large Social Network. The New England Journal of Medicine, 358: 2249 - 2258. DeCicca, P., Kenkel, D., Mathios, A., Shin, Y-J., Lim, J-Y. (2007), Youth Smoking, Cigarette Prices and Anti-Smoking Sentiments. Health Economics, 17: 733 - 749. Ennett, S., Bauman, K., Foshee, V., Pemberton, M., and Hicks, K. (2001), Parent-child communication about adolescent tobacco and alcohol use: What do parents say and does it affect youth behavior?. Journal of Marriage and Family, 63:4862. Eisenberg, D., and Quinn, B. (2006), Estimating the Effect of Smoking Cessation on Weight Gain: An Instrumental Variable Approach. Health Services Research, 41(6): 2255 2266. Fang. H., Ali, MM., Rizzo, JA. (2009), Does Smoking Affect Body Weight and Obesity in China. Economics and Human Biology, 7(3): 334 350. Fletcher J. (2010), Social Interactions and Smoking Decisions: Evidence Using Multiple Cohorts, Instrumental Variables and School Fixed Effects. Health Economics, 19(4): 466 484.. Hall, J.A., and Valente, T.W. (2007), Adolescent smoking networks: The effects of influence and selection on future smoking. Addictive Behaviors, 32(12), 3054- 3059. Hoffman, B.R., Monge, P.R., Chou, C.-P., and Valente, T.W. (2007), Perceived peer influence and peer selection on adolescent smoking. Addictive Behaviors, 32(8), 1546- 1554. Jones, A. (1989), A Double-Hurdle Model of Cigarette Consumption. Journal of Applied Econometrics, 4: 23 39.

  • 23

    K. Kan, W-D Tsai. (2004), Obesity and Risk Knowledge. Journal of Health Economics, 23(5): 907-934. B. Katzman, S. Markowitz, and K A. McGreory. (2007), An Empirical Investigation of the Social Market for Cigarettes. Health Economics, 16: 1025-1039. Lee. S. (2007), Endogeneity in quantile regression models: a control function approach. Journal of Econometrics 141: 1131-1158. K. Lahiri and J. Song. (2000), The effect of smoking on health using a sequential self- selection model. Health Economics, 9:491511. L. Liang, F. Chaloupka, M. Nichter, and R. Clayton. (2001), Prices, policies and youth smoking. Addiction, 98(1):105122. E. Lloyd-Richardson, G. Papandonatos, A. Kazura, C. Stanton, and R. Niaura. (2002), Differentiating stages of smoking intensity among adolescents: Stage-specific psychological and social influences. Journal of Consulting and Clinical Pyschology, 70(4):9981009. Manski, C. (1993), Identification of Endogenous Social Effects: The Reflection Problem. Review of Economic Studies, 60(3): 531 542. Manski, C. (2000), Economic Analysis of Social Interactions. Journal of Economic Perspectives. 14(3):115 136. Mayhew, K., Flay, B., and Mott, J. (2000), Stages in the development of adolescent smoking Drug and Alcohol Dependence, S9(1):S61S81. Newhouse, J., and McClellan, M. (1998), Econometrics in Outcomes Research: The Use of Instrumental Variables. Annual Review of Public Health, 19:1734. Norton, E., Lindrooth, R., and Ennett, S. (1998), Controlling for the endogenity of peer substance use on adolescent alcohol and tobacco use. Health Economics, 7:439 453. L. Powell and F. Chaloupka. (2005), Parents, public policy and youth smoking. Journal of Policy Analysis and Management, 24(1):93112. L. Powell, J. Tauras, and H. Ross. (2005), The importance of peer effects, cigarette prices and tobacco control policies for youth smoking behavior. Journal of Health Economics, 24:950968. Schoenbaum, M., Unutzer, J., McCaffrey, D., Duan, N., Sherbourne, C., and Wells, K. (2002), The Effects of Primary Care Depression Treatment on Patients Clinical Status and Employment. Health Services Research, 37(5):114558.

  • 24

    Terza, JV., Basu, A., Rathouz, PJ. (2008), Two Stage Residual Inclusion Estimation: Addressing Endogenity in Health Econometric Modeling. Journal of Health Economics, 27(3): 531 543. Valente, T., Unger, J., Johnson, A. (2005), Do Popular Students Smoke? The Association among Popularity and Smoking among Middle School Students. Journal of Adolescent Health, 37(4): 323 329.

  • 25

    Table 1: Descriptive Statistics from Wave I (1994)

    Variables Mean Std. Dev. Min Max

    Smoking Stage: - - Never Smoker 0.49 0.49 0 1

    Experimental Smoker 0.25 0.34 0 1 Intermittent Smoker 0.17 0.37 0 1

    Regular Smoker 0.09 0.29 0 1 Tax 0.33 0.16 0.025 0.75

    Policy Index 1.35 0.95 -0.79 1.93 Age 15.15 1.74 11 20

    Grade 9.67 1.63 7 12 Male 0.49 0.50 0 1 White 0.61 0.48 0 1 Black 0.23 0.42 0 1

    Hispanic 0.17 0.37 0 1 Religious 2.73 2.01 1 9

    Parent Smoke 0.29 0.45 0 1 Easy Access to

    Cigarettes 0.31 0.46 0 1

    Lives with both Parents 0.49 0.50 0 1 Mother College 0.25 0.43 0 1 Father College 0.21 0.41 0 1 Work Fulltime 0.29 0.45 0 1

    N: 19,988

  • 26

    Table 2: Factor Analysis Results - Eigenvalues

    Eigenvalue Factor 1 Factor 2 Factor 3 Factor 4 Factor 5 Factor 6 Factor 7 Factor 8

    State Policy 3.47424 0.98259 0.34966 0.13239 -0.05573 -0.10363 -0.16013 -0.21682 Parent-Child Relationship

    2.25837 0.85097 -0.00106 -0.06302 -0.08920 -0.11075 -0.15053 -0.22606

    Table 3: Factor Loadings and Scoring Coefficients State Policies Factor Loadings Scoring Coefficients Factor 1 Factor 1 Vending Machines Banned in location accessible to youths

    0.5769 0.07586

    Billboard Prohibited within 500 feet of school

    0.7056 0.12497

    Free Sample Prohibited 0.0363 -0.01104 Marketing Restrictions 0.8654 0.36874 Local Enforcement of Youth Access

    -0.7305 -0.16956

    Dissemination of Information 0.8028 0.26638 Education Measurement for Compliance

    -0.4308 -0.05026

    Schools Required to Offer Tobacco use Prevention Programs

    -0.7277 -0.13654

    Table 4: Factor Loadings and Scoring Coefficients Parent-Child Relationship

    Factor Loadings Scoring Coefficients Factor 1 Factor 1

    Respondents Survey: Parents Care 0.2117 0.05434

    Parents Understand 0.2168 0.06921 Parents Pay Attention 0.2747 0.09283

    Communicate 0.2728 0.06833 Parents Survey:

    Get Along 0.7941 0.36525 Understand 0.6504 0.19047

    Trust 0.7360 0.26783 Decision 0.6491 0.18838

  • 27

    Table 5: Ordered Probit Estimates (Marginal Effects) without Social Learning Theory

    Variables LR Chi2: 2085.60

    Log Likelihood: -25360.05 N: 19,988

    Never Smoker Experimental Smoker Intermittent Smoker Regular Smoker

    Tax 0.039** (0.019)

    -0.008 (0.010)

    -0.017 (0.020)

    -0.014** (0.007)

    Policy Index 0.017** (0.005)

    0.003 (0.005)

    -0.001 (0.003)

    -0.010** (0.004)

    Parent Characteristics

    Mom College 0.080*** (0.008)

    -0.017 (0.020)

    -0.025 (0.030)

    -0.028*** (0.003)

    Dad College 0.021*** (0.004)

    -0.005 (0.004)

    -0.009 (0.008)

    -0.007*** (0.001)

    Work Fulltime -0.015*** (0.004)

    0.003*** (0.001)

    0.006 (0.004)

    -0.005 (0.005)

    Demographics Age -0.025***

    (0.005) 0.005*** (0.001)

    0.011*** (0.002)

    0.009*** (0.002)

    Male 0.004 (0.006)

    -0.001 (0.001)

    -0.002 (0.003)

    -0.001 (0.002)

    White -0.074*** (0.010)

    0.017*** (0.002)

    0.033*** (0.004)

    0.025*** (0.003)

    Black 0.149*** (0.012)

    -0.041*** (0.004)

    -0.064*** (0.005)

    -0.045*** (0.003)

    Hispanic 0.059*** (0.010)

    -0.015*** (0.003)

    -0.026*** (0.004)

    -0.019*** (0.003)

    Grade -0.020*** (0.005)

    0.004*** (0.001)

    0.009*** (0.002)

    0.007*** (0.002)

    Religious 0.094*** (0.006)

    -0.019*** (0.001)

    -0.042*** (0.003)

    -0.034*** (0.002)

    Notes: Include school fixed effects. Robust Standard Errors in parenthesis. *** - sig. at the 1% level; ** - sig. at the 5% level; * - sig. at the 10% level.

  • 28

    Table 6: Ordered Probit Estimates (Marginal Effects) with Social Learning Theory

    Variables LR Chi2: 6404.44

    Log Likelihood: -23200.63 N: 6,549

    Never Smoker Experimental Smoker Intermittent Smoker

    Regular Smoker

    Tax -0.007

    (0.013) 0.002

    (0.003) 0.003

    (0.007) 0.002

    (0.003) Policy Index 0.005

    (0.004) -0.001 (0.001)

    -0.003 (0.002)

    -0.001 (0.001)

    Peer Smoking Nominated -0.193***

    (0.003) 0.048*** (0.002)

    0.101*** (0.002)

    0.04*** (0.001)

    Parent Characteristics

    Lives with both parents

    0.055*** (0.008)

    -0.014*** (0.002)

    -0.029*** (0.004)

    -0.013*** (0.002)

    Easy Access to Cigarettes

    -0.010*** (0.007)

    0.021*** (0.001)

    0.053*** (0.004)

    0.027*** (0.002)

    Parent Smoke -0.029*** (0.008)

    0.007*** (0.002)

    0.016*** (0.004)

    0.007*** (0.002)

    Parent-Child Relationship Index

    0.036*** (0.004)

    -0.009*** (0.001)

    -0.019*** (0.002)

    -0.009*** (0.001)

    Mom College -0.001 (0.004)

    0.000 (0.001)

    0.000 (0.002)

    0.000 (0.001)

    Dad College 0.003 (0.004)

    -0.000 (0.000)

    -0.000 (0.002)

    -0.000 (0.001)

    Work Fulltime -0.46*** (0.014)

    0.007 (0.010)

    -0.003 (0.019)

    0.017*** (0.007)

    Demographics Age 0.002

    (0.005) -0.001 (0.001)

    -0.001 (0.002)

    -0.001 (0.001)

    Male 0.003 (0.006)

    -0.001 (0.002)

    -0.002 (0.003)

    -0.001 (0.002)

    White -0.073*** (0.010)

    0.019*** (0.003)

    0.038*** (0.005)

    0.017*** (0.002)

    Black 0.084*** (0.012)

    -0.024*** (0.004)

    -0.043*** (0.006)

    -0.018*** (0.002)

    Hispanic 0.030** (0.010)

    -0.008** (0.003)

    -0.016** (0.005)

    -0.007** (0.002)

    Other -0.076** (0.025)

    0.013*** (0.003)

    0.042** (0.014)

    0.022** (0.009)

    Grade -0.032*** (0.005)

    0.008*** (0.001)

    0.017*** (0.003)

    0.008*** (0.001)

    Urban 0.019** (0.007)

    -0.005** (0.002)

    -0.010** (0.004)

    -0.005** (0.002)

    Religious 0.043*** (0.007)

    -0.010*** (0.002)

    -0.022*** (0.003)

    -0.010*** (0.002)

    Notes: Include school fixed effects. Robust Standard Errors in parenthesis. *** - sig. at the 1% level; ** - sig. at the 5% level; * - sig. at the 10% level.

  • 29

    Table 7: Ordered Probit Estimates (Marginal Effects) with Social Learning Theory

    Variables LR Chi2: 2100.93

    Log Likelihood: -18347.56 N: 19,988

    Never Smoker Experimental Smoker

    Intermittent Smoker

    Regular Smoker

    Tax 0.001

    (0.015) -0.002 (0.004)

    -0.004 (0.007)

    -0.002 (0.004)

    Policy Index 0.003 (0.004)

    -0.001 (0.001)

    -0.002 (0.002)

    -0.001 (0.001)

    Peer Smoking Classmates

    -0.519***

    (0.036) 0.130*** (0.010)

    0.240*** (0.017)

    0.149*** (0.011)

    Parent Characteristics

    Lives with both parents

    0.056*** (0.009)

    -0.014*** (0.002)

    -0.026*** (0.004)

    -0.016*** (0.003)

    Easy Access to Cigarettes

    -0.122*** (0.008)

    0.026*** (0.002)

    0.058*** (0.004)

    0.040*** (0.003)

    Parent Smoke -0.046*** (0.010)

    0.011*** (0.002)

    0.022*** (0.005)

    0.014*** (0.003)

    Parent-Child Relationship Index

    0.055*** (0.004)

    -0.014*** (0.001)

    -0.026*** (0.002)

    -0.016*** (0.001)

    Mom College 0.003 (0.016)

    0.006 (0.011)

    -0.019 (0.020)

    0.004 (0.008)

    Dad College 0.021 (0.017)

    -0.022* (0.012)

    0.021 (0.024)

    -0.004 (0.009)

    Work Fulltime -0.046*** (0.014)

    0.007 (0.010)

    -0.003 (0.019)

    0.017** (0.007)

    Demographics Age -0.016**

    (0.006) 0.004** (0.001)

    0.008** (0.003)

    0.005** (0.002)

    Male -0.001 (0.007)

    0.000 (0.002)

    0.001 (0.003)

    0.000 (0.002)

    White -0.089**** (0.012)

    0.023*** (0.003)

    0.041*** (0.005)

    0.025*** (0.003)

    Black 0.117*** (0.014)

    -0.034*** (0.005)

    -0.053*** (0.006)

    -0.030*** (0.003)

    Hispanic 0.032** (0.012)

    -0.009** (0.003)

    -0.015** (0.005)

    -0.009** (0.003)

    Other -0.077** (0.029)

    0.014* (0.004)

    0.037** (0.015)

    0.026** (0.012)

    Grade -0.014** (0.006)

    0.003** (0.001)

    0.006** (0.003)

    0.004** (0.002)

    Urban 0.025** (0.008)

    -0.006** (0.002)

    -0.012** (0.004)

    -0.007** (0.002)

    Religious 0.061*** (0.008)

    -0.014*** (0.002)

    -0.029*** (0.004)

    -0.018*** (0.002)

    Notes: Include school fixed effects. Robust Standard Errors in parenthesis. *** - sig. at the 1% level; ** - sig. at the 5% level; * - sig. at the 10% level.

  • 30

    Table 8: Control Function Ordered Probit Estimates (Marginal Effects)

    Variables LR Chi2: 7509.06 Log Likelihood: -22648.32

    N: 6,549 Never

    Smoker Experimental

    Smoker Intermittent

    Smoker Regular Smoker

    Tax -0.007

    (0.005) 0.004

    (0.015) 0.005

    (0.011) 0.001

    (0.003) Policy Index 0.001

    (0.001) -0.001 (0.004)

    0.005 (0.003)

    -0.004 (0.003)

    Peer Smoking Nominated -0.306***

    (0.018) 0.204* (0.107)

    0.045 (0.041)

    0.057 (0.064)

    Parent Characteristics

    Easy Access to Cigarettes

    -0.078*** (0.001)

    0.028*** (0.009)

    0.022*** (0.007)

    0.028*** (0.003)

    Parent Smoke -0.035*** (0.001)

    0.024** (0.010)

    0.006*** (0.001)

    0.005** (0.002)

    Parent-Child Relationship Index

    0.032*** (0.001)

    -0.004*** (0.001)

    -0.021*** (0.003)

    -0.007*** (0.001)

    Residual 0.492*** (0.087)

    -0.114*** (0.021)

    -0.238*** (0.042)

    -0.145*** (0.025)

    Notes: Estimates of only main variables of interest provided. Include school fixed effects. Robust Standard Errors in parenthesis. *** - sig. at the 1% level; ** - sig. at the 5% level; * - sig. at the 10% level. Instruments: % of peers whose parents smoke; % of peers with easy access to cigarettes at home; % of peers who lives with both parents.

  • 31

    Table 9: Control Function Ordered Probit Estimates (Marginal Effects)

    Variables LR Chi2: 2597.79 Log Likelihood: -18099.13

    N: 19,988 Never

    Smokers Experimental

    Smoker Intermittent

    Smoker Regular Smoker

    Tax -0.006

    (0.005) 0.002

    (0.017) 0.003

    (0.013) 0.001

    (0.005) Policy Index 0.001

    (0.001) 0.002

    (0.005) 0.005

    (0.004) -0.008 (0.005)

    Peer Smoking Classmates -0.735***

    (0.017) 0.227

    (0.405) 0.256*** (0.024)

    0.252*** (0.018)

    Parent Characteristics

    Easy Access to Cigarettes

    -0.094*** (0.001)

    0.021*** (0.010)

    0.031*** (0.008)

    0.042*** (0.004)

    Parent Smoke - 0.039*** (0.003)

    0.010** (0.004)

    0.020** (0.009)

    0.009* (0.004)

    Parent-Child Relationship Index

    0.053** (0.001)

    -0.009*** (0.002)

    -0.030*** (0.004)

    -0.014*** (0.001)

    Residual 0.208*** (0.013)

    -0.048*** (0.004)

    -0.100*** (0.007)

    -0.061*** (0.001)

    Notes: Estimates of only main variables of interest provided. Include school fixed effects. Robust Standard Errors in parenthesis. *** - sig. at the 1% level; ** - sig. at the 5% level; * - sig. at the 10% level. Instruments: % of peers whose parents smoke; % of peers with easy access to cigarettes at home; % of peers who lives with both parents.

  • 32

    Appendix 1: First Stage Estimation (OLS) of Control Function Approach

    Variables Nominated Peers Classmates

    Tax -0.002

    (0.008) -0.001 (0.001)

    Policy Index -0.006 (0.005)

    -0.002 (0.001)

    Instruments Peer Parent Smoke 0.030***

    (0.006) 0.219*** (0.008)

    Peer Easy Access to Cigarettes 0.136*** (0.014)

    0.294*** (0.009)

    Peer Lives with both parents -0.041*** (0.012)

    -0.021*** (0.006)

    Parent Characteristics Lives with both parents -0.014

    (0.011) 0.003

    (0.002) Easy Access to Cigarettes 0.021

    (0.017) 0.006

    (0.002) Parent Smoke 0.026

    (0.018) 0.006

    (0.002) Parent-Child Relationship

    Index -0.030***

    (0.014) -0.001 (0.001)

    Mom College 0.006 (0.013)

    0.003 (0.002)

    Dad College 0.004 (0.012)

    0.005** (0.002)

    Work Fulltime 0.028*** (0.012)

    0.006 (0.004)

    Demographics Age 0.011

    (0.008) 0.006*** (0.001)

    Male 0.020*** (0.008)

    0.001 (0.002)

    White 0.074*** (0.016)

    0.015*** (0.002)

    Black -0.090*** (0.019)

    -0.036 (0.023)

    Hispanic -0.011 (0.015)

    -0.046 (0.026)

    Other 0.021 (0.046)

    0.023 (0.036)

    Grade 0.023*** (0.009)

    0.033*** (0.001)

    Urban -0.005 (0.003)

    -0.006 (0.004)

    Religious -0.010 (0.004)

    -0.012 (0.009)

    Adjusted R2 0.085 0.407 F Statistic 31.20 690.55

    N 6,549 19,988 Notes: Include school fixed effects. Robust Standard Errors in parenthesis. *** - sig. at the 1% level; ** -

    sig. at the 5% level; * - sig. at the 10% level.