Att. B Q&Q Justification

Attachment B_QxQ justification.docx

Child Support Noncustodial Parent Employment Demonstration (CSPED)

Att. B Q&Q Justification

OMB: 0970-0439

Document [docx]
Download: docx | pdf



ATTACHMENT B:

question-by-question justification

AND SURVEYS REFERENCED

FOR THE CSPED Baseline survey






Table B.1. Baseline Child Support 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 demographic and socioeconomic characteristics. Thus, these variables are important for use as covariates and to form subgroups. In addition, these variables will be used to describe the characteristics of the population served by CSPED and to predict participation in CSPED services.

Country of birth (B3)

BSF

X

X

X

X


Age arrived in US (B4)

BSF

X

X




Marital Status (B5)

OMB tailored for CSPED






Highest level of education (B6)

CBRA, tailored for CSPED

X

X

X

X


Biological Child Roster

Number of biological children (C1)

PACT

X

X


X


The number of biological children will be used to predict participation in CSPED services and to determine the number of children about whom the information in the subsequent child-specific questions of this survey will be collected.

Child’s name
(C2-C4)

BSF






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

Child’s gender (C6)

PACT

X

X

X

X


Research has shown that parents’ behaviors are associated with the gender of their children. For example, the birth of a son increases fathers’ labor supply and wages more than the birth of a daughter (Lundberg and Rose 2002), so fathers’ investments in children appear to be larger for sons than daughters. For never-married mothers, the birth of a son is associated with higher increases in the speed of marriages to the child’s father than it is the birth of a daughter (Lundberg and Rose 2003). 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). However, other studies find the opposite or no association between child gender and father involvement (Seltzer 1991).

Child’s date of birth or age (C7–C8)

PACT

X

X

X

X


This demographic information can be used to confirm the child’s identity at study follow-ups. Research has also found 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 or father’s name
(C9-C11)

PACT

X

X




This information will be used to identify the child’s mother or father to fill in the mother’s or father’s name in later survey questions. Identifying the mother or father of each child is particularly important, since it allows us to identify noncustodial parents who experience multiple partner fertility, which is associated with diminished noncustodial parent contact with the child (Manning and Smock 2000; Manning, Stewart, and Smock 2003). Multiple partner fertility will be used to describe the characteristics of study participants and as a covariate in the impact analysis.

Whether mother and father were married or living together when child born (C13 and C14)

BSF

X

X




Nonresident fathers who were married to the child’s mother at the time of the child’s birth are more likely to be involved with their children than other nonresident fathers (Seltzer 1991).

Paternity was voluntarily acknowledged (C15)


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 be predictive of subsequent relationships between co-parents and between fathers and their children.

Court established paternity (C16)

BSF

X

X


X


Whether ever lived with child (C17)

FFCWS tailored for CSPED

X

X

X

X


For parents who live with their children, some amount of parent-child contact is almost inevitable, while nonresidential parents are likely to have little or no contact with their children (Minton and Pasley 1996, Seltzer 1991). Therefore, we expect that the noncustodial parent’s baseline residence status will be a strong predictor of involvement with the child at follow-up. Furthermore, residential status at baseline is a useful variable for defining subgroups. For example, to examine whether child support programs had larger effects on residential or nonresidential parents or affected different outcomes for residential versus nonresidential parents.

Number of nights stayed with child in last month (C18–C19)

FFCWS tailored for CSPED; WFNJ tailored for CSPED

X

X

X

X


Who has primary responsibility for child (C20a-C20b)

FFCWS tailored for CSPED

X

X

X



Any contact with child in last month (C21)

EHS tailored for CSPED

X

X

X

X

X

A key goal of parenting programs is to increase parental involvement with children. The extent of parent-child interaction at baseline is expected to be an important predictor of subsequent parental involvement. In addition, other studies have found that parent-child baseline contact is a useful variable to create subgroups for analysis. For instance, an evaluation of the Parents’ Fair Share program found larger impacts on father involvement in sites with the lowest levels of baseline contact (Miller and Knox 2001).

In-person contact with child in last month (C22)

EHS tailored for CSPED

X

X

X

X

X

Assessment of relationship quality (C23)

EHS tailored for CSPED

X

X

X

X

X

Relationships

Status of relationship with mother/father (D2–D3)


BSF tailored for CSPED

X

X

X

X

X

Child support programs may improve noncustodial parents’ relationship skills and co-parenting relationships, increasing the likelihood that parents are involved with the other parent of at least one of their children at follow-up. Including measures of noncustodial parents’ relationships with the other parent at baseline in the analyses will increase the precision of the estimates of the impact of child support programs on parents’ 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), and children born to married parents may experience higher levels of parental involvement than children born to non-married parents (Selzer, 1988). Thus, a noncustodial parent’s relationship with the other parent at baseline could also be an important covariate in estimating impacts on noncustodial parent involvement. Finally, program impacts might differ by relationship status and relationships status might predict program participation.

Lives with mother/father (D4)

BSF tailored for CSPED

X

X




Noncustodial parent’s baseline residence status will be used as a predictor of involvement with the child at follow-up and to describe the characteristics of the participants in the study at baseline.

Nights in past 30 days stays with mother/father (D5)

PACT tailored for CSPED

X

X




Whether custodial mother/father has a romantic partner she/he lives with (D6)

WFNJ tailored for CSPED

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 noncustodial parent involvement at follow-up.

Quality of relationship with mother/father (D7)

FFCWS

X

X



X

Child support programs might improve relationships between noncustodial parents and the other parent of their children. Relationship quality may also be predictive of father engagement (Fagan and Palkovitz 2011).

Quality of the collaborative co-parenting relationship (D8)

PAM

X

X


X

X

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

Whether child support order is in place (D9)

BSF tailored for CSPED

X

X


X

X

A key goal of CSPED is to promote responsible parenting, including noncustodial parents’ material support of their children. Financial support of children through formal and informal monetary payments and in-kind purchases will be important measures of CSPED’s impact. By including a measure of financial support at baseline in the impact analyses, 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 (D10–D16)

BSF, FFCWS tailored for CSPED

X

X


X

X

Barriers to parent involvement (D17-D18)

EHS tailored for CSPED

X

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 parent-child distance will be a useful covariate. CSPED might have smaller impacts on noncustodial parent involvement when parents live a larger distance from their child. In addition, barriers to noncustodial parent involvement might also be barriers to program participation.

Whether parent has other romantic partner (D19-D20)


PACT, WFNJ tailedored for CSPED

X

X


X


Having a spouse or a cohabiting partner is associated with more favorable labor market outcomes for men (Cohen 2002, Cornwell and Rupert 1997, Nock 1998) and economic benefits for women (Light, 2002), so noncustodial parents’ relationships with new partners are expected to be predictive of later economic outcomes.

Having a new partner 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).

Finally, relationships between noncustodial parents and new partners might also predict program participation.

Parents’ relationship status with partner (D21)

BSF tailored for CSPED

X

X


X


Whether parent lives with partner (D22)

BSF tailored for CSPED

X

X


X


Number of nights parent spent with partner in past 30 days (D23)

CSPED-developed

X

X


X


Whether partner has children under age 18 (D24)

PACT tailored for CSPED

X

X


X


Noncustodial parent involvement with the children of a new partner might be a predictor of current and future involvement of the noncustodial parent with his/her own children, and it can also be a predictor of program participation.

Whether partner’s children stayed with partner and respondent in last 30 days (D25)

PACT tailored for CSPED

X

X


X


Economic Stability

Paid work in last month (E1)

WFNJ tailored for CSPED

X

X

X

X

X

A key goal of child support programs is to improve noncustodial parents’ economic self-sufficiency. Noncustodial parents’ current employment status, earnings, and barriers to employment are expected to be key predictors of similar economic outcomes at follow-up. It is also possible that the effects of CSPED on noncustodial parents’ economic outcomes will vary according to a parent’s baseline labor market experience. For example, the Parents’ Fair Share program increased earnings only among men with the least labor market experience (Miller and Knox 2001). Employment, earnings, health insurance coverage and use of public benefits might also be related to program participation.

Date of last employment (E2)


WFNJ tailored for CSPED

X

X




Earnings and hours worked in last month (E3–E9)


WFNJ tailored for CSPED, RWTW tailored for CSPED

X

X


X

X

Barriers to employment (E10)

FFCWS, BSF tailored for CSPED

X

X


X

X

Food stamp benefits received (E11)

ACS tailored for CSPED

X

X


X

X

Health insurance coverage (E12)

ACS tailored for CSPED

X

X


X

X

Rent or own home (E13)

WFNJ

X

X


X


Housing instability, including homelessness, eviction, frequent moves, 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). Housing instability is especially prevalent amongst low-income families with children, and women who have been incarcerated experience greater difficulties securing employment and housing than those who have not (Phinney et al, 2007). Understanding the housing circumstances of the CSPED sample will help capture the extent of their disadvantages. Housing instability has also been cited as a barrier to employment (Miller and Knox 2001), and so can be an important covariate in models of CSPED’s impact on noncustodial parents economic well-being. Finally, housing instability might predict program participation.

Co-residence with parents or grandparents (E14-E15)

CSPED-developed

X

X





Anticipated housing stability (E16-E17)

HII

X

X


X


Father/Mother Background and Well-Being

Co-residence with own parents during childhood (F1)

FFCWS

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 for both men and women.

Own father/mother’s involvement in childhood (F2)

PACT

X

X




Quality of relationship with own father/mother (F3)

PACT

X

X




Depressive symptoms (F4)

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, which includes eight of the nine items from the PHQ-9; has been shown to be a useful measure of depression in population-based studies (Kroenke et al. 2009).

Parental stress and self-assessment as a parent (F5-F5b)

PSI

X

X

X

X

X

Parental stress is an indicator of parents’ own well-being and is also correlated with parental engagement and the quality of the co-parenting relationship (Bronte-Tinkew, Horowitz, and Carrano 2010). Thus, this measure of parenting stress will be used as both an outcome and a useful covariate for increasing the precision of other impact estimates. Additionally, whether the noncustodial parent has experienced aggravation in the parenting role will be a useful variable to create subgroups, as those parents who have been aggravated 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 can be a predictor of program participation.

Locus of control and future orientation (F6)


FFCWStailored for CSPED, PACT

X

X


x

X

Disadvantaged noncustodial parents may feel helpless to change their circumstances and pessimistic about the future. If programs help noncustodial parents 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.

Ever convicted of a crime (F7)

SVORI

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 CSPED 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 face unique challenges in securing work and housing (Geller and Curtis 2011, Pager 2003), as do previously incarcerated women (Phinney et al, 2007). Criminal history information can be used as covariates in the impact analyses to improve the precision of our impact estimates. Finally, parole or probation status might predict program participation.

Longest/most recent incarceration (F8-F10)

SVORI tailored for CSPED

X

X

X



Current parole or probation (F11)

SVORI tailored for CSPED

X

X


X


Motivation to Participate in Program

Motivators to apply to program (G1)


PACT tailored for CSPED

X



X


Participation is a common challenge in programs serving low-income 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.

Importance of program participation to respondent (G2)

PACT tailored for CSPED

X



X


Contact Information

A1–7

H1–11k







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


Sources: Parents and Children Together (PACT), 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), Youthbuild, the American Community Survey (ACS).

SURVEYS REFERENCED

The list below contains brief descriptions of the eight surveys referenced in the CSPED 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 CSPED baseline survey’s goals.

1. Parents and Children Together (PACT)

Mathematica’s OMB-approved Parents and Children Together (PACT) impact and evaluation study assesses innovative approaches to helping fathers increase involvement in the lives of their children and achieve economic stability. Similarly to the PACT study, the CSPED study uses a random assignment design to examine the effects of parenthood and employment services provided to low-income parents. Thus, the CSPED data collection instruments were designed to draw heavily on the PACT study instruments, which will facilitate comparisons of program outcomes between the two studies. .

Intake processes for the CSPED study are also closely modeled after the PACT study. In addition, the CSPED and PACT studies use an MIS to perform random assignment and to track program participation, and both studies include qualitative interviews with program staff, a web survey of staff and community partners, baseline data collection with parents via telephone survey, and a 12-month follow-up survey on various outcomes related to family and economic well-being.

The PACT baseline data collection instrument served as the starting point for the CSPED baseline data collection instrument. The CSPED evaluation team reviewed each question within the PACT instrument and made modifications. These modifications fall into three general categories:

  • Minor wording modifications The CSPED target population varies slightly from the PACT sample population in that noncustodial mothers are included in the CSPED study. For this reason, gendered pronouns and question wording were modified throughout the instrument to accommodate noncustodial mothers in addition to fathers. Other minor wording changes were also made to reflect programmatic variations, areas of analytical focus, clarify target behaviors, and maximize the reliability and validity of data collected from the target CSPED population.

  • Deletion of items excluded from analysis In order to reduce respondent burden, the CSPED evaluation team removed any baseline items that would not be used for analysis of the CSPED baseline survey. Examples include items about respondent religiosity, country of origin and disability status.

  • Addition of items required for analysis Several items were added in order to better understand program effects on participant outcomes. Examples include questions about employer-provided health insurance coverage, additional detail about the respondent’s living situation, and a self-assessment of the respondent as a parent.

Surveys are available from Mathematica upon request.



2. 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.

3. 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.

4. 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

5. 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

6. 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.

7. 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.

8. 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.

9. YouthBuild

Mathematica’s YouthBuild study is an evaluation of youth and community development programs targeted towards out-of-school youth from low-income families. The evaluation measures core program outcomes, including educational attainment, postsecondary education planning, employment, earnings, delinquency, involvement with the criminal justice system, and social and emotional development. Mathematica, as a subcontractor to MDRC, is designing and implementing a web survey of YouthBuild grantees and three mixed-mode surveys of youth that will take place 12, 30 and 48 months after random assignments.

Surveys are available from Mathematica upon request.

10. Parenting Alliance Measure (PAM)

The PAM is a parental assessment designed to provide assessment of parental perceptions of the strength of their parenting alliance. The PAM is a screening tool suitable for family counselors, joint custody evaluations, identification of issues with parenting skills, and is also used to assess the impact of intervention programs.

Assessments are available for purchase from PAR, Inc.

11. Patient Health Questionnaire Screeners (PHQ)

The PHQ is a clinical tool designed to provide clinicians with screening and diagnostic tools for mental health disorders. All PHQ instruments have been tested in clinical settings, and are designed to improve recognition rates of depression and anxiety.

Assessments are available from Pfizer at www.phqscreeners.com.

12. The American Community Survey (ACS)

The ACS is an ongoing survey of American households. It is administered annually using a multi-mode design and collects demographic, employment, disability, health, and spending data in order to inform federal and state funding decisions.

Surveys are available from the United States Census.

13. Housing Instability Index (HII)

The HII is a tool created for the Safe Housing Assistance with Rent Evaluation (SHARE) study, a CDC-funded evaluation designed to learn about the connection between domestic violence and housing. The HII provides information about vulnerability, quality of life and health outcomes associated with housing stability.

The index is available through the National Alliance to End Homelessness.

14. The Parenting Stress Index (PSI)

The PSI is designed to identify dysfunctional parenting and identify sources of stress within a family unit. This empirically validated measure has been tested across languages and cultures and can be used as a diagnostic and predictive tool for future parental behavior problems.

Assessments are available for purchase from PAR, Inc.



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.

Light, Audrey. “Gender Differences in the Marriage and Cohabitation Income Premium.” Demography, vol. 41, no. 2, 2004, pp. 263-284.

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.

Phinney, Robin, Sheldon Danziger, Harold Pollack, and Kristen Seefeldt. “Housing Instability Among Current and Former Welfare Recipients.” Americal Journal of Public Health, vol. 95, no. 5, 2007, pp. 832-837.

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. “Children’s Contact with Absent Parents.” Journal of Marriage and the Family, (1988): 663-677.

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
File TitleAttachment B Question by Question Justification
AuthorSheena Flowers
File Modified0000-00-00
File Created2021-01-27

© 2024 OMB.report | Privacy Policy