PACT Appendix C - RF Baseline Q-by-Q Justification - 10-31-12 (2)

PACT Appendix C - RF Baseline Q-by-Q Justification - 10-31-12 (2) (2).docx

Parents and Children Together (PACT) Evaluation

PACT Appendix C - RF Baseline Q-by-Q Justification - 10-31-12 (2) (2)

OMB: 0970-0403

Document [docx]
Download: docx | pdf



Parents and Children Together (PACT) Evaluation


APPENDIX C


question-by-question justification,


KEY OUTCOME domains AND SUBGROUPS,


AND SURVEYS REFERENCED FOR


THE PACT RESPONSIBLE FATHERHOOD Baseline survey




INCLUDED IN PREVIOUSLY-APPROVED ICR: OCTOBER 31, 2012






Table C.1. Baseline RF Survey: Question-by-Question Justification

Question

Source

How Question Will be Used

Rationale

Descriptor

Covariate

Subgroup

Predictor of Participation

Outcome

Introduction

Introduction (i1–7)







Obtaining consent

Demographic and Socioeconomic Characteristics

Race and ethnicity (B1–B2)

OMB

X

X

X



Program impacts may be moderated by these variables and thus they are important for use as covariates and subgroups. In addition, they will be used to describe the characteristics of the population served by PACT and to predict participation.

Country of birth (B3)

BSF

X

X

X

X


Age arrived in US (B4)

BSF

X

X




Highest level of education (B5)

CBRA tailored for PACT

X

X

X

X


Education completed in US (B5a)

PACT-developed

X

X




Biological Child Roster

Any biological children (C1)

PACT-developed






Only individuals who have at least one biological child will be eligible for PACT.

Number of biological children (C2)

PACT-developed

X

X


X


The number of biological children will be used to establish how many children to collect information on in the subsequent child-specific questions and to predict participation.

Child’s name
(C3a-c)

BSF






This information will be used to identify the child for subsequent follow-ups and to fill in the child’s name in later survey questions.

Child’s gender (C5)


X

X

X

X


In general, fathers’ investments in children appear to be larger for sons than daughters. The birth of a son increases fathers’ labor supply and wages more than the birth of a daughter (Lundberg and Rose 2002). For never-married mothers, the birth of a son increases the speed of marriages to the child’s father more than the birth of a daughter (Lundberg and Rose 2003), and among parents married at the time of the child’s birth, fathers are more likely to live with sons than daughters at the child’s first birthday (Lundberg, McLanahan, and Rose 2007). Some studies suggest that adolescent girls receive less attention from fathers than do sons (Harris and Morgan 1991) and that nonresident fathers’ involvement with girls is more likely to decline over time (Manning and Smock 1999), although others find the opposite or no association between child gender and father involvement (Seltzer 1991).

Child’s date of birth or age (C6–C7)


X

X

X

X


This demographic information can be used to confirm the child’s identity at follow-up. There is also mixed evidence that father involvement can vary with the child’s age (Hofferth et al. 2002, Seltzer 1991, Veum 1993), suggesting that child age will be a useful covariate.

Mother’s name
(C8-C9a)


X

X




This information will be used to identify the child’s mother for subsequent follow-ups and to fill in the mother’s name in later survey questions. Identifying the mother of each child is particularly important, since it allows us to identify fathers who experience multiple partner fertility, which is associated with diminished father-child contact (Manning and Smock 2000; Manning, Stewart, and Smock 2003). Multiple partner fertility will be used as a descriptor and covariate in the impact analysis.

Whether parents were married when child born (C11)

BSF

X

X




On average, nonresident fathers who were married to the child’s mother at the time of the child’s birth have more involvement with their children than other nonresident fathers (Seltzer 1991).

Paternity voluntarily acknowledged (C12)


BSF

X

X


X


Establishing paternity is an important step in ensuring that unwed fathers provide financial support for their children. We expect that legal establishment of paternity will be a strong predictor of father involvement and especially of fathers’ financial support of children, since paternity establishment triggers child support orders. Voluntary paternity establishment may be an indicator of the father’s desire to be involved with and assume responsibility for the child, and may also be associated with the quality of the parents’ relationship at the time of the child’s birth. We expect that this measure will therefore be predictive of subsequent relationships between co-parents and between fathers and their children.

Court established paternity (C13)

BSF

X

X


X


Who child lives with (C14)

FFCWS tailored for PACT

X

X

X

X


For fathers who live with their children, some amount of father-child contact is almost inevitable, while a nontrivial share of nonresidential fathers have little or no contact with their children (Minton and Pasley 1996, Seltzer 1991). Therefore, we expect that the father’s baseline residence status will be a strong predictor of his involvement with the child at follow-up. Furthermore, we may be interested in examining whether RF programs had larger effects on residential or nonresidential fathers or affected different outcomes for residential versus nonresidential fathers. Residential status at baseline is therefore useful for defining subgroups.

Number of nights stayed with child in last month (C15–C15a)

FFCWS tailored for PACT, WFNJ tailored for PACT

X

X

X

X


Whether ever lived with child (C16)

FFCWS

X

X

X



Time spent with child in last month (C17)

BSF, EHS

X

X

X

X

X

A key goal of RF programs is to increase fathers’ involvement with their children. The extent of father-child interaction at baseline is expected to be an important predictor of subsequent father involvement. Given that Parents’ Fair Share found larger impacts on father involvement in sites with the lowest levels of baseline contact, father-child baseline contact may also be a useful subgroup for analysis (Miller and Knox 2001).

Other contact with child in last month (C18)

EHS tailored for PACT

X

X

X

X

X

In-kind material support of child in the last month (C19–C19a)

BSF tailored for PACT, BSF

X

X



X

A key goal of PACT is to promote responsible parenting, including fathers’ material support of their children. By measuring financial support at baseline, we will improve the precision of our impact estimates.

Barriers to Father Involvement

Distance from child is a barrier (C20)

EHS

X

X

X

X


The geographic distance between nonresident fathers and their children is negatively associated with father involvement (Manning and Smock 1999, Seltzer 1991, Veum 1993), so father-child distance will be a useful covariate. We might also expect that PACT will have smaller impacts on father involvement when fathers live a larger distance from their child. Finally, barriers to father involvement might also be barriers to program participation.

Child’s mother is a barrier (C21)

EHS tailored for PACT

X

X


X


Mothers play an important role in facilitating or impeding fathers’ involvement with their children, including when parents live apart (Fagan and Barnett 2003). We propose to include the father’s perception that the mother is a barrier to his involvement with the child as a covariate in the impact analysis. Barriers to father involvement might also be barriers to program participation.

Another family member is a barrier (C22)

EHS tailored for PACT

X

X


X


Other family members may also facilitate or impede fathers’ involvement with their children; for example, better relationships with mothers’ extended families are predictive of greater nonresident father involvement (Ryan, Kalil, and Ziol-Guest 2008). We propose to include the father’s perception that other family members are a barrier to his involvement with the child as a covariate in the impact analysis. Barriers to father involvement might also be barriers to program participation.

Financial difficulties are a barrier (C23)

EHS tailored for PACT

X

X


X


Low-income fathers report that financial constraints can be a barrier to their involvement with their children, either because they lack the money to pay for activities with their children or because guilt over their inability to provide financially for their children diminishes their contact with the child (Johnson and Doolittle 1998; Nelson, Clampet-Lundquist, and Edin 2002). Barriers to father involvement might also be barriers to program participation.

Relationships

Status of relationship with mother (D2–D3)


BSF tailored for PACT, BSF

X

X

X

X

X

RF programs may improve fathers’ relationship skills and co-parenting relationships, increasing the likelihood that fathers are involved with the mother of at least one of their children at follow-up. Measuring fathers’ relationships with mothers at baseline will increase the precision of the estimates of the impact of RF programs on fathers’ romantic relationships. A father’s romantic relationship with the child’s mother may also be a key predictor of his contact with his children (Tach, Mincy, and Edin 2010), so a father’s relationship with the mother at baseline may also be an important covariate in estimating impacts on father involvement. Finally, program impacts might differ by relationship status and relationships status might predict program participation.

Contact with mother (D4)

PACT-developed

X

X




Among romantically involved fathers, fathers that live with mothers are more likely to be involved with their children than fathers who do not (Johnson 2001); therefore, living arrangements may be an important covariate in estimating impacts on father involvement.


Lives with mother (D5)

BSF tailored for PACT

X

X




Nights per week stays with mother (D6)

PACT-developed

X

X




Whether mother has a romantic partner she lives with (D7)

WFNJ

X

X




Fathers’ involvement with their nonresidential children is significantly lower when mothers are involved with new partners (Guzzo 2009; Tach, Mincy, and Edin 2010). We propose to collect this information at baseline to improve the precision of our estimates of father involvement at follow-up.

Quality of relationship with mother (D8)

FFCWS

X

X



X

RF programs might improve relationships between fathers and the mothers of their children. Relationship quality may also be predictive of father engagement (Fagan and Palkovitz 2011).

Quality of the collaborative co-parenting relationship (D9)

PAM

X

X


X

X

The quality of the co-parenting relationship is predictive of subsequent father involvement (Carlson, McLanahan, and Brooks-Gunn 2008; Sobolewski and King 2005). Measuring the quality of co-parenting at baseline will improve the precision of our impact estimates, both for subsequent co-parenting and other outcomes of interest. Finally, the quality of the co-parenting relationship might predict program participation.

Child’s mother supports father in how he wants to raise child (D10)

FFCWS tailored for PACT

X

X


X

X

Whether child support order in place (D11)

BSF tailored for PACT

X

X


X

X

A key goal of PACT is to promote responsible parenting, including fathers’ material support of their children. Financial support of children through formal and informal monetary payments and in-kind purchases will be important measures of PACT’s impact. By measuring financial support at baseline, we will improve the precision of our impact estimates. Finally, child support status might predict program participation.

Formal and informal support paid to mother in last month (D12–D14)

BSF, FFCWS

X

X


X

X

Whether father has other romantic partner (D15–D16)


PACT-developed, WFNJ

X

X


X


Spouses and cohabiting partners are associated with more favorable labor market outcomes for men (Cohen 2002, Cornwell and Rupert 1997, Nock 1998), so fathers’ relationships with new partners are expected to be predictive of later economic outcomes.

Fathers’ new partners may also be associated with less involvement by fathers in the lives of their children from prior relationships (Carlson, McLanahan, and Brooks-Gunn 2008; Manning and Smock 1999; Seltzer 1991; Tach, Mincy, and Edin 2010).

Relationships with new partners might also predict program participation.

Father’s relationship status with partner (D17–D18)

BSF tailored for PACT, BSF

X

X


X


Whether father lives with partner (D19)

BSF tailored for PACT

X

X


X


Whether partner makes it hard to see child (D20)

EHS tailored for PACT

X

X


X


Nonbiological, Residential Children

Nonbiological children in household (E1)

PACT-developed

X

X

X

X


Residential, nonbiological children may divert fathers’ resources from their biological children, particularly if they do not live with their biological children. We also are interested in whether this association varies with whether men consider themselves as “social fathers” to nonbiological children. Manning and Smock (1999) find a small negative association between the number of co-resident stepchildren and father visitation to nonresidential biological children.

Number of nonbiological children living in household (E2)

PACT-developed

X

X




Number of nonbiological children living in household that belong to current partner (E3)

PACT-developed

X

X




Economic Stability

Paid work in last month (F1)

WFNJ tailored for PACT

X

X

X

X

X

A key goal of RF programs is to improve fathers’ economic self-sufficiency. Fathers’ current employment status and earnings are expected to be key predictors of similar economic outcomes at follow-up. It is also possible that the effects of PACT on fathers’ economic outcomes will vary according to a man’s baseline labor market experience. For example, Parents’ Fair Share increased earnings only among men with the least labor market experience (Miller and Knox 2001). Employment and earnings might also be related to program participation.

Date of last employment (F2)


WFNJ tailored for PACT

X

X




Earnings in last month (F3–F5)


WFNJ tailored for PACT, RWTW tailored for PACT

X

X


X

X

Rent or own home (F6)

WFNJ

X

X


X


Housing instability, including homelessness, eviction, frequent movies, involuntary moves due to being unable to pay rent or mortgage, and living with others without paying rent, is experienced by a considerable share of urban men, especially those who have been incarcerated (Geller and Curtis 2011). Understanding the housing circumstances of the PACT sample will help capture the extent of the disadvantage. Housing instability has also been cited as a barrier to employment (Miller and Knox 2001), and so may be an important covariate in models of PACT’s impact on fathers’ economic well-being. Finally, housing instability might predict program participation.

Anticipated housing stability (F7)

HII

X

X


X


Father Background and Well-Being

Own father’s involvement in childhood (G1)

PACT-developed

X

X




Men’s relationships with their own fathers are associated with their understanding of the fatherhood role (Forste, Bartkowski, and Jackson 2009; Roy 2006). We propose to include these measures as covariates in the impact models.

Quality of relationship with father (G2)

PACT-developed

X

X




Depressive symptoms (G3)

PHQ-8

X

X


X

X

Parental depression has been shown to have adverse consequences for child outcomes (Downey and Coyne 1990, Gelfand and Teti 1990). To measure depressive symptoms, we will use eight items from the Patient Health Questionnaire (PHQ-9), which was designed as a diagnostic instrument for depression but can also be used to measure subthreshold depressive disorder in the general population (Martin et al. 2006). The PHQ-9 has been shown to be reliable and valid in diverse populations and has been used in clinical settings to measure symptom improvement and monitor treatment outcomes (Kroenke, Spitzer, and Williams 2001; Löwe et al. 2004). Findings from telephone administrations of the instrument have been shown to be similar to in-person assessments (Pinto-Meza et al. 2005). The PHQ-8 includes eight of the nine items from the PHQ-9; it has been shown to be a useful measure of depression in population-based studies (Kroenke et al. 2009).

Parental stress (G4)

PSI

X

X

X

X

X

Parental stress is an indicator of parents’ own well-being and is also correlated with father engagement and the quality of the coparenting relationship (Bronte-Tinkew, Horowitz, and Carrano 2010). Thus, fathers’ parenting stress will be both an outcome and a useful covariate for increasing the precision of other impact estimates. Additionally, fathers who experience aggravation in the parenting role may be a useful subgroup, as these fathers may both be more motivated to participate in order to improve their relationships and may have the potential for greater improvements in outcomes at follow-up. Finally, parental stress may be a predictor of program participation.

Locus of control and future orientation (G5)


FFCWS, PACT-developed

X

X


x

X

Disadvantaged fathers may feel helpless to change their circumstances and pessimistic about the future. If RF programs help fathers acquire new skills or improve their circumstances, through employment for example, they may develop greater feelings of self-efficacy and the ability to plan for the future.

Religious service attendance (G6)

BSF

X

X




Frequency of religious service attendance is positively associated with mental health outcomes and negatively associated with distress (Ellison et al. 2001) and marital disruption (Call and Heaton 1997). Therefore, religious attendance is a useful covariate in models estimating the impact of PACT on fathers’ well-being and romantic relationships.

Ever arrested (G7)


BSF

X

X




Recent research suggests that a history of incarceration and involvement with the criminal justice system may be fairly common among fathers in the PACT target population (Pettit and Western 2004). Parental incarceration has major negative effects on child and family well-being, reducing the financial support and other types of support parents can provide to their children and families. Previously-incarcerated men also face unique challenges in securing work and housing (Geller and Curtis 2011, Pager 2003). Criminal history information can be used to improve the precision of our impact estimates. Finally, parole or probation status might predict program participation.

How many times arrested (G8–G8a)


SVORI, SVORI tailored for PACT

X

X




Number of times convicted of a crime (G9–G9a)

SVORI

X

X




Longest incarceration (G10)

SVORI tailored for PACT

X

X

X



Current parole or probation (G11)

SVORI tailored for PACT

X

X


X


Motivation to Participate in Program

Area of life most want to change (H1)


PACT-developed

X



X


Participation is a common challenge in programs serving low-income fathers and couples (Avellar et al. 2011; Dion et al. 2010). Past research has shown that factors such as motivation to change and perceived benefits of services are associated with subsequent participation (Dumas et al. 2007, Eisner and Meidert 2011, Nock et al. 2006, Nock et al. 2007). We will collect this information to estimate the impact on those who receive services as well as the impact of being offered services.

Barriers to participation (H2)

PACT-developed

X



X


Contact Information

A1–8

I1–10







Contact information is necessary to locate the respondent for the follow-up data collection 12 months later.


Sources: Building Strong Families Study (BSF), American Recovery and Reinvestment Act COBRA Subsidy Study (CBRA), Fragile Families and Child Well-Being Study (FFCWS), Early Head Start Research and Evaluation Project (EHS), Work First New Jersey (WFNJ), Parenting Alliance Measure (PAM), Rural Welfare-to-Work Demonstration Evaluation (RWTW), Housing Instability Index (HII), Patient Health Questionnaire (PHQ-9), Parenting Stress Index (PSI), Serious Violent Offender Reentry Initiative Evaluation (SVORI).


KEY OUTCOME DOMAINS



Father involvement

Coparenting

Parenting

Attitudes, knowledge, and skills

Marriage and romantic relationships

Employment

Mental health

Service receipt

Child outcomes

KEY SUBGROUPS


Race/ethnicity

Educational attainment

Child’s gender

Child’s age

Residential status at baseline

Father involvement at baseline

Relationship with child’s mother at baseline

Employment status at baseline



SURVEYS REFERENCED

The list below contains brief descriptions of the seven surveys referenced in the PACT baseline survey, as well as locations of the surveys referenced. Descriptions were compiled from websites about the surveys and descriptions of Mathematica studies were gathered from project summaries. When necessary, we modified questions drawn from these surveys to make them easier to understand or to have the questions align more closely with the baseline survey’s goals.

1. Building Strong Families Study (BSF)

The United States Department of Health and Human Services/Administration for Children and Families (ACF) initiated the Building Strong Families (BSF) project to help interested and romantically involved low-income, unwed parents build stronger relationships and thus enhance their child’s well being and their own future. The BSF evaluation being conducted by Mathematica is designed to test the effectiveness of these programs for couples and children. BSF data collection included a baseline information form to collect demographic and socioeconomic data along with two follow-up surveys. The follow-up surveys included questions related to mother-father relationships, family structure, fathers’ involvement in child rearing, parent-child relationships and the home environment, family functioning, child well-being and development, and parental well-being.

Surveys are available from Mathematica upon request.

2. American Recovery and Reinvestment Act COBRA Subsidy Study (CBRA)

Sponsored by the U.S. Department of Labor, Mathematica’s American Recovery and Reinvestment Act (ARRA) COBRA Subsidy study examines the effect of the availability of an ARRA COBRA premium subsidy on the take-up of COBRA coverage and other health and employment outcomes. As part of the study, Mathematica will conduct a survey of COBRA-eligible individuals drawn from state Unemployment Insurance recipients. The CBRA survey asks questions related to respondents’ demographic characteristics, employment history, receipt of social services, and health insurance.

Surveys are available from Mathematica upon request.

3. Fragile Families and Child Well-Being Study (FFCWS)

The Fragile Families and Child Well-Being Study is a longitudinal study of a cohort of nearly 5,000 children born between 1998 and 2000 from birth through age five. Approximately one-third of the children were born to unmarried parents. Interviews were conducted with both mothers and fathers covering a range of topics including attitudes, relationships, and parenting behavior.

Study protocols and codebooks can be found here: http://www.fragilefamilies.princeton.edu/documentation.asp

4. Early Head Start Research and Evaluation Project (EHS)

The Early Head Start Research and Evaluation Project was a national, large-scale, random assignment evaluation of Early Head Start. The study included interviews with both mothers and fathers about child and family functioning when children were 14 months through 36 months of age. One branch of the study focused on low-income fathers’ involvement in their children’s lives.

Study protocols can be found here: http://www.acf.hhs.gov/programs/opre/ehs/ehs_resrch/index.html

5. Work First New Jersey (WFNJ)

Mathematica evaluated the effects of New Jersey’s initiative to help welfare recipients transition from welfare to work. WFNJ interviewed sample members annually for five years documenting changes in household composition, income, employment, and other indicators of well-being.

Surveys are available from Mathematica upon request.

6. Rural Welfare-to-Work Demonstration Evaluation (RWTW)

Mathematica’s Rural Welfare-to-Work Strategies Demonstration Evaluation used random assignment to assess innovative approaches to helping welfare-dependent and other low-income families in rural areas to enter, maintain, and advance in employment and to secure family well-being. Data collection included a baseline information form to collect demographic and socioeconomic data on sample members and two follow-up surveys to collect detailed employment history data as well as information on various outcomes related to individual and family well-being.

Surveys are available from Mathematica upon request.

7. Evaluation of the Serious Violent Offender Reentry Initiative (SVORI)

The Evaluation of the Serious Violent Offender Reentry Initiative (SVORI) was a multi-year, multi-site evaluation funded by National Institute of Justice. The impact evaluation was designed to measure the impact of enhanced reentry programming on post-release outcomes. As part of the evaluation, interviews were conducted at four points in time.

Surveys are available from the National Archive of Criminal Justice Data.



REFERENCES

Avellar, Sarah M., Robin Dion, Andrew Clarkwest, Heather Zaveri, Subuhi Asheer, Kelley Borradaile, Megan Hague Angus, Timothy Novak, Julie Redline, and Marykate Zukiewicz. “Catalog of Research: Programs for Low-Income Fathers.” Washington, DC: Mathematica Policy Research, June 2011.

Bronte-Tinkew, J., Horowitz, A., and Carrano, J. “Aggravation and Stress in Parenting: Associations with Coparenting and Father Engagement Among Resident Fathers.” Journal of Family Issues, vol. 31, 2010, pp. 525–555.

Call, V.R.A., and T.B. Heaton. “Religious Influence on Marital Stability.” Journal for the Scientific Study of Religion, vol. 36, 1997, pp. 382–392.

Carlson, M. J., S.S. McLanahan, and J. Brooks-Gunn. “Coparenting and Nonresident Fathers’ Involvement with Young Children After a Nonmarital Birth.” Demography , vol. 45, 2008, pp. 461–88.

Cohen, P.N. “Cohabitation and the Declining Marriage Premium for Men.” Work and Occupations, vol. 29, 2002, pp. 346–363.

Cornwell, C., and P. Rupert. “Unobservable Individual Effects, Marriage and the Earnings of Young Men.” Economic Inquiry, vol. 35, 1997, pp. 285–294.

Dion, Robin, Sarah M. Avellar, and Elizabeth Clary. “Implementation of Eight Programs to Strengthen Unmarried Parent Families.” Washington, DC: Mathematica Policy Research, May 2010.

Downey G., and J.C. Coyne. “Children of Depressed Parents: An Integrative Review.” Psychological Bulletin, vol. 108, 1990, pp. 50–76.

Dumas, J.E., Nissley-Tsiopinis, J., and Moreland, A.D. “From Intent to Enrollment, Attendance, and Participation in Preventative Parenting Groups.” Journal of Child and Family Studies, vol. 16., no. 1, 2007, pp. 1–26.

Eisner, M. and U. Meidert. “Stages of Parental Engagement in a Universal Parent Training Program.” Journal of Primary Prevention, vol. 34, 2011, pp. 951–963.

Ellison, C.G., J.D. Boardman, D.R. Williams, and J.S. Jackson. “Religious Involvement, Stress, and Mental Health: Findings from the 1995 Detroit Area Study.” Social Forces, vol. 80, 2001, pp. 215–249.

Fagan, J., and M. Barnett. “The Relationship Between Maternal Gatekeeping, Paternal Competence, Mothers’ Attitudes About the Father Role, and Father Involvement.” Journal of Family Issues, vol. 24, 2003, pp. 1020–1043.

Fagan, J., and R. Palkovitz. “Coparenting and Relationship Quality Effects on Father Engagement: Variations by Residence, Romance.” Journal of Marriage and Family, vol. 73, 2011, pp. 637-653.

Forste, R., J.P. Bartkowski, and R.A. Jackson. “‘Just Be There for Them’: Perceptions of Fathering Among Single, Low-Income Men.” Fathering, vol. 7, 2009, pp. 49–69.

Gelfand, D.M., and D.M. Teti. “The Effects of Maternal Depression on Children.” Clinical Psychology Reviews, vol. 10, 1990, pp. 329–353.

Geller, A., and M.A. Curtis. “A Sort of Homecoming: Incarceration and the Housing Security of Urban Men.” Social Science Research, vol. 40, 2011, pp. 1196–1213.

Guzzo, K.B. “Maternal Relationships and Nonresidential Father Visitation of Children Born Outside of Marriage.” Journal of Marriage and Family, vol. 71, 2009, pp. 632–649.

Harris, K. M., and S.P. Morgan. “Fathers, Sons, and Daughters: Differential Paternal Involvement in Parenting.” Journal of Marriage and the Family, vol. 53, 1991, pp. 531–544.

Hofferth, S.L., J. Pleck,J.L. Stueve, S. Bianchi, and L. Sayer. “The Demography of Fathers: What Fathers Do.” In Handbook of Father Involvement: Multidisciplinary Perspectives, edited by C.S. Tamis-LeMonda and N. Cabrera. Mahwah, NJ: Lawrence Erlbaum Associates, 2002.

Johnson, E.S., and F. Doolittle. “Low-Income Parents and the Parents’ Fair Share Program: An Early Qualitative Look at Improving the Ability and Desire of Low-Income Noncustodial Parents to Pay Child Support.” In Fathers Under Fire: The Revolution in Child Support Enforcement, edited by I. Garfinkel, S. McLanahan, D.R. Meyer, and J.A. Seltzer. New York: Russell Sage Foundation, 1998.

Johnson, Waldo E., Jr. “Paternal Involvement Among Unwed Fathers.” Children and Youth Services Review, vol. 23, 2001, pp. 513-536.

Kroenke, Kurt, Robert L. Spitzer, and Janet B.W. Williams. “The PHQ-9: Validity of a Brief Depression Severity Measure.” Journal of General Internal Medicine, vol. 16, no. 9, 2001, pp. 606–613.

Kroenke, Kurt, Tara W. Strine, Robert L. Spitzer, Janet B.W. Williams, Joyce T. Berry, and Ali H. Mokdad. “The PHQ-8 as a Measure of Current Depression in the General Population.” Journal of Affective Disorders, vol. 144, no. 1, 2009, pp. 163–173.

Löwe, B., J. Unützer, C.M. Callahan, A.J. Perkins, and K. Kroenke, “Monitoring Depression Treatment Outcomes with the Patient Health Questionnaire-9.” Medical Care, vol. 42, no. 12, 2004, pp. 1194–1201.

Lundberg, S., S. McLanahan, and E. Rose. “Child Gender and Father Involvement in Fragile Families.” Demography, vol. 44, 2007, pp. 79–92.

Lundberg, S., and E. Rose. “The Effect of Sons and Daughters on Men’s Labor Supply and Wages.” Review of Economics and Statistics, vol. 84, 2002, pp. 251–268.

Lundberg, S., and E. Rose. “Child Gender and the Transition to Marriage.” Demography, vol. 40, 2003, pp. 333–349.

Manning, W.D., and P.J. Smock. “New Families and Nonresident Father-Child Visitation.” Social Forces, vol. 78, 1999, pp. 87–116.

Manning, W.D., and P.J. Smock. “‘Swapping’ Families: Serial Parenting and Economic Support for Children.” Journal of Marriage and Family, vol. 62, 2000, pp. 111–122.

Manning, W.D., S.D. Stewart, and P.J. Smock. “The Complexity of Fathers’ Parenting Responsibilities and Involvement with Nonresident Children.” Journal of Family Issues, vol. 24, 2003, pp. 645–667.

Martin, A., W. Reif, A. Klaiberg, and E. Braehler, “Validity of the Brief Patient Health Questionnaire Mood Scale (PHQ-9) in the General Population.” General Hospital Psychiatry, vol. 28, no. 1, 2006, pp. 71–77.

Miller, C., and V. Knox. “The Challenge of Helping Low-Income Fathers Support Their Children.” Manpower Demonstration Research Corporation, November 2001.

Minton, C., and K. Pasley. “Fathers’ Parenting Role Identity and Father Involvement: A Comparison of Nondivorced and Divorced, Nonresident Fathers.” Journal of Family Issues, vol. 16, 1996, pp. 26–45.

Nelson, T.J., S. Clampet-Lundquist, and K. Edin. “Father Involvement Among Low-Income, Noncustodial African-American Fathers in Philadelphia.” In Handbook of Father Involvement: Multidisciplinary Perspectives, edited by C.S. Tamis-LeMonda, and N. Cabrera. Mahwah, NJ: Lawrence Erlbaum Associates, 2002.

Nock, M.K., Ferriter, C., and Holmberg, E. “Parent Beliefs about Treatment Credibility and Effectiveness: Assessment and Retention of Subsequent Treatment Participation.” Journal of Child and Family Studies, vol. 16, 2007, pp. 27–38.

Nock, M.K., and V. Photos. “Parent Motivation to Participate in Treatment: Assessment and Prediction of Subsequent Participation.” Journal of Child and Family Studies, vol. 15, no. 3, 2006, pp. 345–358.

Nock, S.L. Marriage in Men’s Lives. New York: Oxford University Press, 1998.

Pager, D. “The Mark of a Criminal Record.” American Journal of Sociology, vol. 108, 2003, pp. 937–975.

Pettit, B. and B. Western. “Mass Imprisonment and the Life Course: Race and Class Inequality in U.S. Incarceration.” American Sociological Review, vol. 69, 2004, pp. 151–169.

Pinto-Meza A., A. Serrano-Blanco, M.T. Peñarrubia, E. Blanco, and J.M. Haro, “Assessing Depression in Primary Care with the PHQ-9: Can It Be Carried Out over the Telephone?” Journal of General Internal Medicine, vol. 20, no. 9, 2005, pp. 738–742.

Roy, K.M. “Father Stories: A Life Course Examination of Paternal Identity Among Low-Income African American Men.” Journal of Family Issues, vol. 27, 2006, pp. 31–54.

Ryan, R., Kalil, A., and Ziol-Guest, “Longitudinal Patterns of Non-Resident Fathers’ Involvement: The Role of Resources and Relations.” Journal of Marriage and Family, vol. 70, 2008, pp. 962–977.

Seltzer, J.A. “Relationships Between Fathers and Children Who Live Apart: The Father’s Role After Separation.” Journal of Marriage and the Family, vol. 53, 1991, pp. 79–101.

Sobolewski, J.M., and V. King. “The Importance of the Coparental Relationship for Nonresident Fathers’ Ties to Children.” Journal of Marriage and Family, vol. 67, 2005, pp. 1196–1212.

Tach, L., R. Mincy, and K. Edin. “Parenting as a ‘Package Deal’: Relationships, Fertility, and Nonresident Father Involvement Among Unmarried Parents.” Demography, vol. 47, 2010, pp. 181–204.

Veum, J.R. “The Relationship Between Child Support and Visitation: Evidence from Longitudinal Data.” Social Science Research, vol. 22, 1993, pp. 229–244.



File Typeapplication/vnd.openxmlformats-officedocument.wordprocessingml.document
AuthorSheena Flowers
File Modified0000-00-00
File Created2021-01-28

© 2024 OMB.report | Privacy Policy