Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Journal of Research in Special Educational Needs doi: 10.1111/1471-3802.12505  Volume   Number   2021 – A systematic review of intervention studies for young children with emotional and behavioral disorders: identifying the research base John William McKenna1 , Frederick Brigham2, Justin Garwood3, Lindsay Zurawski2, Murat Koc1, Carlos Lavin4 and Robai Werunga1 1 4 University of Massachusetts Lowell, USA; 2George Mason University, USA; 3University of Vermont, USA; College Of Charleston, USA Key words: Emotional and/or behavioural disorders, emotional disturbance, behaviour disorder, early childhood special education, evidence-based practice, systematic review. Schools continue to experience difficulty meeting the needs of young children who are formally identified as having an emotional and/or behavioural disorder (EBD). Although schools are mandated to use evidence-based practices to improve student outcomes, such practices must first be identified before they can be employed. Systematic reviews of intervention studies are commonly performed to identify evidence-based practices, make recommendations for service delivery, and identify areas for future research that are needed to inform practice. At this time, researchers have yet to perform a systematic review of intervention studies involving young children identified with EBD – early childhood (EC) – grade 2. The purpose of this study was to identify and describe published school-based intervention research for this student population. Studies meeting selection criteria were evaluated according to the relevant What Works Clearinghouse (WWC) Design standards to identify studies with strong internal validity that reported positive effects. Twenty-nine manuscripts reporting 30 intervention studies were identified. Findings suggest that practitioners must primarily rely on their professional judgement and values guided by principles embedded in their training when planning instruction and support for young children with EBD due to the absence of high-quality intervention research. Additional rigorous evaluations are needed so that practice is better informed by science. Recommendations for practice, areas for future research, and study limitations are discussed. An increasing number of young children exhibit high levels of challenging behaviour (Carter et al., 2010). Compounding this issue, behavioural difficulties that manifest at a young age may continue as children grow older (Bulotsky-Shearer et al., 2010), placing them at increased risk for social and school difficulties (Garwood ª 2021 NASEN et al., 2017b). In sum, a failure to provide timely and effective intervention to young children with problem behaviours has deleterious effects (see Conroy et al., 2015). In regard to serving young children who display characteristics consistent with an emotional and/or behavioural disorder (EBD), Child Find (i.e. identification) continues to be elusive (see Kauffman and Landrum, 2013; Mitchell, Kern, and Conroy, 2019). Identification of young children with EBD represents a significant challenge to the field, as identification is a prerequisite to the provision of services and supports that are sufficiently intensive and comprehensive (Conroy and Brown, 2004). Identification of young children with EBD Symptomology of EBD includes externalizing (e.g. aggression and non-compliance) and internalizing (anxiety and depression) behaviours (Kauffman and Landrum, 2013). Formal identification of EBD may be delayed for as many as 6–20 years despite characteristics consistent with the disorder being present during the early years. A number of factors confound timely and accurate identification. First, the behaviour of young children with EBD may be misinterpreted as age appropriate or normative (Wakschlag et al., 2007). For example, a failure to identify young children with EBD may occur when temper tantrums, defiance and aggression are perceived as normative or typical for young children (Withey, 2018). Second, the federal definition for Emotional Disturbance (ED), the special education disability category that is consistent with many characteristics of EBD, has little utility for serving young children with EBD because it does not specifically address their characteristics (Conroy and Brown, 2004). For example, performing problem behaviours for a ‘long period of time’ is a defining characteristic of ED. This time requirement may inadvertently prevent young children from being formally identified as a child in need and eligible for special education services. The federal definition of ED also requires adverse effects on academic achievement (Garwood et al., 2020). These 1 Journal of Research in Special Educational Needs,  – criteria may also inadvertently prevent the identification of young children as early childhood (EC) programming may not be academic based. Furthermore, the academic performance of young children with ED may not be sufficiently different from that of their peers without ED to be of immediate concern. Third, stigma associated with a diagnosis consistent EBD or ED may also contribute to the under or delayed identification of children (Farmer, 2013). Fourth, children who perform behaviours consistent with an internalising disorder may be particularly difficult to identify (Poulou, 2015). These factors provide context to the limited number of young children nationally who are formally identified with ED. Furthermore, the characteristics and, thus, needs of children who are formally identified with EBD are likely to be different from those considered at risk because the characteristics of ‘at risk’ lacks precision and vary across research teams (see Maag, 2006; Williams et al., 2016). Thus, inferences drawn from research focusing on students considered at risk for EBD may not be salient to children who display characteristics consistent with EBD. Therefore, it is critical that young children with EBD receive research-based intervention that targets their unique needs to interrupt the progressively negative effects of this disorder and improve short- and long-term outcomes. The identification rate of students with EBD (i.e. those who receive special education services under the category of ED) is less than 1% of the school-age population (U.S. Department of Education, 2017), but this number greatly underestimates the prevalence of EBD nationally. Forness et al. (2012) provided two different types of estimates to depict a more accurate picture of prevalence rates for EBD. Point prevalence (PP) indicates the percentage of all school-aged students exhibiting EBD at one particular time. Cumulative prevalence (CP), which may be more accurate than PP because it accounts for fluctuation in the expression of EBD across time (e.g. expression of EBD characteristics may be episodic), indicates the percentage of students who would meet the criteria for EBD at any point in their schooling. Conservative estimates suggest PP at approximately 12% and CP at approximately 37% (Forness et al., 2012a; Forness et al., 2012b). In sum, data suggests that many students with EBD do not receive special education services despite displaying characteristics with the disorder. Importance of research-based instruction and intervention Young children formally identified with EBD require timely and effective intervention to prevent or ameliorate the negative effects of EBD (Garwood et al., 2017a; Mitchell et al., 2019). As such, federal policies (e.g. ESSA, 2015; IDEIA, 2004) emphasize the use of research-based practices to improve student outcomes. Yet, educators of young children report infrequent use of effective practices for preventing and responding to challenging behaviours (Hoover et al., 2012; Vinh et al., 2016). However, before practitioners can employ research-based practices, an accumulation of rigorous (i.e. allowing some degree of causal inference) intervention research is necessary to identify the practices that are most likely to be effective (see McKenna et al., 2019). During the 2016–2017 academic year, only 2680 students between the ages of three and five received special education services for ED (U.S. Department of Education, 2017). This number represents 0.35% of all children in this age range who were identified with a disability. The number of children between the ages of five and eight who were formally identified with ED is unavailable, as federal child count data is not disaggregated in this manner. Research does suggest that students with EBD are not typically identified for supplemental supports (i.e. special education) until later in their schooling, such as fourth grade (Malmgren and Meisel, 2002) or middle school (Kauffman and Landrum, 2013). However, the needs of young children (e.g. EC to 2nd grade) who are formally identified may be different (e.g. more intense, pervasive and resistant to intervention) than children at risk for EBD/ED because their behaviours are likely far from what is considered normative (e.g. represent a significant level of impairment), resulting in identification despite the aforementioned challenges to Child Find (see Costello, Egger, and Angold, 2005; Forness et al., 2012a; Kulkarni and Sullivan, 2019; The Peacock Hill Working Group, 1991; Zabel, Kaff, and Teagarden, 2013). 2 Importance of systematic reviews Systematic reviews of the literature are commonly performed to report existing research, identify effective practices and make recommendations for future investigations to inform school practice (Cook et al., 2016b; Cook et al., 2016a; Maggin et al., 2013). Systematic reviews are common in the field of EBD, as evidenced by studies investigating topics such as social studies content (Garwood et al., 2019) and cross-age peer tutoring (Watts et al., 2019) interventions. Furthermore, systematic reviews may apply expert panel recommendations to evaluate extant research and identify investigations with sufficient rigor to make some degree of causal inference between the use of interventions and changes in dependent variables. Through this type of evaluation, research teams can make recommendations based on high-quality studies. At this time, we are unaware of any systematic effort to identify, describe and evaluate intervention studies targeting young children (e.g. EC to grade 2) formally identified with EBD. Study purpose In this systematic review, we sought to determine the extent to which researchers have investigated the effects of interventions on the behavioural, social and academic performance of young children with EBD. Furthermore, we sought to identify promising practices for practitioner ª 2021 NASEN Journal of Research in Special Educational Needs,  – use, as indicated by a quality indicator coding. Considering the necessity of providing timely and effective intervention to young children with EBD, a systematic review in this area is warranted. This investigation was informed by the following research questions: 1. What are the characteristics (e.g. designs, participants, settings and independent variables) of studies investigating the effects of interventions for young children with EBD? 2. According to a What Works Clearinghouse (WWC) design standards evaluation, what interventions may be promising for improving outcomes for young children with EBD? Method First, we conducted two electronic searches of the literature to identify intervention studies involving young children with EBD. Both searches were conducted from the years 1990 to 2018 using ERIC, PsycINFO, Education Research Complete and Academic Search Premier. A start date of 1990 was chosen because this is the year that IDEA was created. The first search used the Boolean phrase ‘behaviour disorder’ AND ‘early childhood’, which yielded 2399 articles. The second search used the Boolean phrase ‘behaviour disorder’ AND ‘intervention’, which yielded 19 345 articles. Titles and abstracts for each article in each of the electronic searches were read to determine suitability for inclusion. Articles meeting initial screening were then read in their entirety to determine if they met the following article selection criteria: (1) the article was published in a peer-reviewed journal between the years 1990 and 2018. We focused specifically on peer-reviewed studies because the process is the ‘gatekeeper’ for special education (see Mitchell et al., 2017); (2) the article was an intervention study that used a single case, comparison group or regression-based design; case studies and descriptive studies were excluded; (3) the study investigated the effects of an intervention delivered in an EC or school setting in the United States; we excluded international studies because other countries may use different criteria for diagnoses and special education services than in the United States (e.g. similar to our exclusion of studies focusing on students considered at risk for EBD, we sought to limit potential variability in student characteristics); (4) the article included at least one participant in EC – grade 2 who received services for ED, EBD or BD, or were diagnosed with the equivalent (e.g. disruptive behaviour disorder, oppositional defiant disorder and anxiety disorder). We included studies that had students with attention deficit hyperactivity disorder (ADHD), attention deficit disorder (ADD), intellectual disability (ID) and Autism as long as the study included at least one additional participant who met selection criteria. We did not include studies with participants who only had a primary disability of ADHD or ADD as part of our selection criteria because these students receive special education services for Other Health Impairment (OHI) ª 2021 NASEN rather than ED; (5) the article included at least one child or student level dependent variable; studies did not need to disaggregate data for students meeting selection criteria to be included in this review; however, disaggregated data was necessary for inclusion in the WWC quality indicator analysis (see below). The first electronic search yielded 3 articles meeting selection criteria. The second electronic search yielded 22 articles with 23 eligible intervention studies. This resulted in total of 26 intervention studies from 25 articles meeting inclusion criteria. Articles were commonly excluded for not including at least one participant that met selection criteria and for not reporting findings from an intervention study. Next, an electronic hand search from the years 2000 to 2018 was completed for the following journals: Behavioral Disorders, Education and Treatment of Children, Journal of Early Intervention, Journal of Emotional and Behavioral Disorders, Preventing School Failure and Topics in Early Childhood Special Education. These journals were selected due to their professional standing and their tendency to publish intervention studies focusing on students with ED and/or young children. Two additional intervention studies meeting selection criteria were identified in the electronic hand search, resulting in a total of 28 intervention studies from 27 articles. Lastly, an additional electronic search of ERIC, PsycINFO, Education Research Complete and Academic Search Premier of the years 1990 to 2018 was performed using the following Boolean phrase: behaviour disorder or emotional disturbance or serious emotional disturbance or emotional and behavioural disorders AND early childhood or young children or pre-school or elementary school. This electronic search was performed in an effort to identify any relevant intervention studies that were not identified in the previous electronic database and hand search. This search yielded 49 252 articles, two of which were a previously unidentified intervention study meeting selection criteria. Upon completion of this process, 30 intervention studies from 29 articles were identified. Figure 1 provides a summary of the procedures used to identify studies meeting selection criteria. Intervention studies meeting selection criteria were independently double coded for descriptive information including participant, setting and intervention characteristics, study design, dependent variables, fidelity data, study results and findings and social validity data. Prior to coding, researchers were trained in the study’s purpose, research questions and the use of an Excel coding sheet to extract relevant data. Researchers then coded one article meeting article selection criteria and disagreements between coders were discussed until 100% agreement was obtained. Initial disagreements tended to focus on intervention characteristics and social validity data. Coders were considered trained and eligible to independently code upon completion of this process. Initial reliability for descriptive coding was 94.8%, with disagreements discussed until 100% agreement was obtained. 3 Journal of Research in Special Educational Needs,  – Figure 1: Search procedure flowchart illustrating search terms, sources and yields for three electronic searches and one hand search of the relevant literature A quality indicator coding using the WWC standards (Kratochwill et al., 2010) was also performed. Studies were first screened by two researchers with previous experience applying the WWC standards to systematic reviews to determine if they employed an eligible design (e.g. comparison group and single case) and reported disaggregated outcomes for children with EBD in EC to grade 2. We focused specifically on studies with disaggregated outcomes because we sought to identify interventions that were specifically effective for young children with EBD. Each study with an eligible design was then independently double coded according to the relevant design standards. Initial reliability on screening was 100%. Initial reliability for design standards coding was 98.6%, with all areas of disagreement discussed until 100% agreement was obtained. Results RQ1: Study characteristics Table 1 reports participant demographic information for each intervention study that met selection criteria. Information is provided on the full participant sample 4 in each study. Table 2 reports information on study characteristics including design, setting, interventions employed, target area of interventions, reported outcomes and fidelity data. Participant characteristics. A total of 442 participants were included in the 30 intervention studies. Of the 442 participants, 73 (16.5%) met participant selection criteria for this review. Thirty-nine (8.8%) had EBD, 12 (2.7%) ED and 11 BD (2.5%). Two (.4%) participants were identified with each of the following disabilities: SED, ODD and ADHD, and emotional handicap and LD. One participant (.2%) was found with each of the following disabilities: Bipolar, BD and ADHD, EBD and ADHD, EBD and language impairment, and EBD, ADD and anxiety disorder. Three participants included in the total for EBD may also have had a secondary disability, but this could not be definitively determined from the article’s information (McDaniel et al., 2016). For six studies (20%), we were unable to determine the exact number of participants meeting selection criteria based on the information provided in the manuscript (Benner et al., ª 2021 NASEN Journal of Research in Special Educational Needs,  – Table 1: Participant demographic information Author (Year) Adkins and Gavins n 3 Age 7–9 Grade 2–3 (2012) Alter et al. (2011) Gender M (2), F Ethnicity Primary disability category SES status AA (3) EBD NR AA (3) EBD NR AA (1), C BD NR ED & ARED NR NR (1) 3 6–9 1–4 M (2), F (1) Beck et al. (2009) 2 6–9 K-3 M (2) (1) Benner et al. (2012) 70 5–10 K-3 M (58), NR F (12) Blair et al. (2000) 1 4 Pre-K M (1) NR BD Clair et al. (2018) 4 8 2 3(M), F AA (4) ID, none (2), Bipolar (1) Cochran et al. (1993) 16 7–11 2–5 M (16) AA (16) BD NR Dawson et al. (2000) 4 7–8 1–2 NR NR EBD NR Dunlap et al. (1996) 3 7–9 2–4 M (1), F NR SED NR (2) Dwyer et al. (2012) 3 7–8 N/A M (3) NR ED NR Falk and Wehby (2001) 6 5–6 K M (6) NR SLD (4), ED/OHI/ADHD (2) NR Hagan-Burke et al. (2015) 1 7 1 M (1) H (1) BD & ADHD NR 7 1–2 M (3) NR EBD NR Mean: 1–3 M (97), W (88), LD (53), MMR (23), EBD (31), PD NR Study 1 Jolivette et al. (2001) Kam et al. (2004) 3 133 8-8 F (36) AA (27), or HI (21), MH (5) O (18) Kennedy et al. (2014) Lopata (2003) 8 24 7–11 6–9 2–4 K-3 M (7), F AA (5), C (1) (3) M (22), H (1), C F (2) EBD NR EBD All on free breakfast and lunch programme (11), AA (12) Malmgren et al. (2005) 3 N/A K, 3, M (3) AA (1), C Mason and Shriner (2008) 6 8–12 2–5 M (5), F AA (2), C McDaniel et al. (2017) 5 7–9 2&3 M (5) McDaniel et al. (2016) 22 7–10 2–4 M (19), &5 EBD NR EBD & ARED NR AA (5) ADHD, ADHD & ODD (3), non- NR AA (14), C EBD NR NR EBD NR H (83%) ADHD, ODD, & CD Avg. family household (2) (1) (4) categorical IEP F (3) (4), H (3), Mixed Race (1) Meyer (1999) 4 N/A 1&3 M (3), F (1) Miller et al. (2014) 11 7–11 N/A M (10), F (1) Peltier and Vannest 4 N/A 2 M (4) income $15k-$59k AA (4) EBD NR (2018) Table 1: (Continued) ª 2021 NASEN 5 Journal of Research in Special Educational Needs,  – Table 1: (Continued) Author (Year) Rafferty (2012) n 4 Age 7–8 Grade 2 Gender 2M, 2F Ethnicity Primary disability category AA (2), H SES status EBD NR NR LD (35), BD (12) NR (1), C (1) Scruggs and Osguthorpe 47 N/A 1–6 (1986) Study 1 Scruggs and Osguthorpe M (30), F (17) 31 N/A 2–5 NR NR LD (24), BD (7) NR 3 7, 8 1, 3, M (3) AA (3) ED (3), Comorbid LD (1) NR M (5), F From EBD (5), EBD & AUT (1) NR EBD (4) OHI (2), LD (1), MMR (1) NR OHI (2), ED (1), SLI (1) NR (1986) Study 2 Trussel et al. (2008) &5 and 11 Weeden et al. (2016) 6 6–9 1–3 (1) minority groups (3) Wehby et al. (2003) 8 7–10 2–4 M (8) AA (6), C Wehby et al. (2005) 4 5–6 K M (4) AA (2), C (2) (2) Notes: SES = socioeconomic status; F = female; M = male; K = kindergarten; NR = not reported; AA = African American; C = Caucasian; H = Hispanic; O = other ethnic origin; W = White; ADHD = Attention Deficit Hyperactivity Disorder; ARED = At Risk for Emotional Disturbance; AUT = Autism; BD = Behavioural Disorder; CD = Conduct Disorder; EBD = Emotional/Behavioural Disorder; ED = Emotional Disturbance; HI = Health Impairment; LD = Learning Disabilities; MH = Multiple Handicaps; MMR = Mild Mental Retardation; ID = Intellectual Disability; ODD = Oppositional Defiant Disorder; OHI = Other health impairment; PD = Physical Disabilities; SED = Severely Emotionally Disturbed; SLD = Speech/language disorder; SLI = Speech/language Impaired. 2012; Kam et al., 2004; Miller et al., 2014; Scruggs and Osguthorpe, 1986, study 1 and 2; Weeden et al., 2016). Only eleven studies (36.6%) included at least one participant meeting selection criteria who were (1) six years of age or younger or (2) in grade one or earlier (Alter et al., 2011; Beck et al., 2009; Blair et al., 2000; Dawson et al., 2000; Falk and Wehby, 2001; Lopata, 2003; Malmgren et al., 2005; Meyer, 1999; Trussel et al., 2008; Weeden et al., 2016; Wehby et al., 2005). Twenty-two studies (73.3%) reported information on participant gender. Of the 73 participants meeting selection criteria, gender was reported for 47 (63.4%). Forty-one (87.2% of participants meeting selection criteria with gender reported) were male. Fifteen studies (50%) reported the ethnicity of participants meeting selection criteria. Ethnicity was reported for 31 of 73 participants (42.5%) meeting criteria. Twenty-one (28.7%) were African American, seven (9.6%) Caucasian, 28.7% (n = 21) and three (4.1%) Hispanic. Information on ethnicity could not be obtained for 42 (57.5%) participants meeting selection criteria (e.g. children with EBD in EC to grade 2). Two studies (6.9%) provided specific information on the socioeconomic status of participants meeting selection criteria. This information was available for 20.1% (n = 22) 6 of the relevant participant sample. In Lopata (2003), 18 students with EBD were eligible for free breakfast and lunch. Six students were in each of the following grades: kindergarten, first and second. In Peltier and Vannest (2018), four second grade students with EBD were eligible for free or reduced-price lunch. Setting characteristics. The majority of studies (n = 23; 76.6%) were conducted in dedicated (i.e. specialised) settings. Nine studies (30%) were conducted in selfcontained classes, seven (23.3%) in self-contained schools, four (13.3%) in general education classrooms, three (10%) in special education classrooms and two (6.6%) in residential schools. One (3.4%) study was conducted in each of the following settings: pre-school setting, resource room, inclusion support classroom, summer treatment programme and summer reading programme. Intervention characteristics. In regards to intervention target area, seventeen studies (56.6%) focused on student behaviour, eight (26.6%) on academics, and five (16.6%) academics and behaviour. Of the eleven studies (36.6%) with at least one relevant participant no older than six years of age or in first grade, seven (63.6%) investigated the effects of behavioural interventions: pre-teaching of reading skills (Beck et al., 2009), adjusting the ratio of toys to children and teaching, prompting and reinforcing ª 2021 NASEN ª 2021 NASEN Table 2: Intervention characteristics, outcomes and fidelity Author (Year) Adkins and Gavins (2012) Design MBD Setting Selfcontained Intervention Self-Regulated Target Academic Reported outcomes • Strategy All participants improved writ- Fidelity data • ing performance. 95% of steps completed across all lessons. Development • (SRSD) Quality of instruction across lessons rated high on a 5-point Likert scale. Alter et al. (2011). MBD Self-contained Teacher-implemented mathematics Academic and • Behaviour Improved problem-solving accu- • Treatment fidelity scores ranged between 75% and racy and on-task behaviour. problem-solving 100% (MN = 97%). strategy • Treatment fidelity checklist also used. Beck et al. (2009) AT General education Pre-teaching reading Behaviour • skills Pre-teaching of unknown items • 25% of intervention yielded higher level of on-task sessions monitored by behaviour for both participants. trained observer using an checklist. • 100% of the items were implemented accurately. Benner et al. (2012) RCT General education Primary-level Behaviour • standard-protocol five-component Intervention significantly • The mean total fidelity decreased problem behaviours. score for the teachers: Increased level of on-task beha- 88% in the treatment behaviour viours was noted but insignifi- group, 6% for the teachers intervention. cant. in the control school. • Table 2: (Continued) 7 Journal of Research in Special Educational Needs,  – implementation Author (Year) Blair et al. (2000) Design AT Setting Pre-school Intervention Antecedent-based Target Behaviour Reported outcomes • reduced intervention • Fidelity data Higher ratio of toys to peers aggression Reversal General education Positive Plus Behaviour • Programme the treatment fidelity, and increased on-task behaviours. sessions were implemented Social skill training led to near- with high degree of accu- zero level of aggression and racy. high level of on-task behaviour Clair et al. (2018) • • QE Self-contained school Cross-age tutoring in reading score was more than able toys to peers. 95% across sessions. Intervention was effective at improving Academic & • Behaviour Procedural reliability regardless of the ratio of avail- off-task • Overall, improvement in sight- Treatment integrity data was high, but no formal behaviour data was reported and academic engagement Cochran et al. (1993) Checklist used to assess • NR • NR • NR word acquisition and use of cooperative statements for tutors and tutees. Dawson et al. (2000) AT Resource room Teacher reading Academic • Reading performance improved model and when reading modelled by tea- computer reading cher or computer. model • Teacher reading model more effective than computer model. Dunlap et al. (1996) Reversal Self-contained Multi-component, Behaviour • assessment-based curricular Improved task engagement, work productivity and accuracy. • Decreased problem behaviour. modification Table 2: (Continued) Journal of Research in Special Educational Needs,  – 8 Table 2: (Continued) ª 2021 NASEN ª 2021 NASEN Table 2: (Continued) Author (Year) Dwyer et al. (2012) Design AT with MBD Setting Self-contained Intervention Replacement of off- Target Behaviour Reported outcomes • Fidelity data Lower levels of off-task and task behaviours in higher levels of replacement behaviour behaviours observed in help, replacement break and choice conditions. • conditions (i.e. • NR • Implementation accuracy Choice resulted in the largest reduction of off-task behaviour. none, help, break or choice between help and break) Falk and Wehby (2001) MBD Self-contained Kindergarten peer- Academic • assisted learning K-PALS led to improved read- levels were 89% for ing performance. strategies (KPALS) decoding, 92% for sound play and 100% for testing. Hagan-Burke et al. (2015) Study AT 1 Special education classroom Function-based Behaviour • Improved task engagement • intervention Whole group choral 100%. response with a slow • MBD with reversal Self-contained Choice – making opportunity Average overall IOA 96% Academic & • Behaviour Choice-making increased appro- • Treatment fidelity was priate and decreased inappropri- 100% across all students ate in each condition. task-related and social behaviour for two of the three participants. Kam et al. (2004) RCT Residential school PATHS (Promoting Behaviour • Reduced rate of growth of tea- • Coder reliability assessed Alternative cher-reported internalising and for 20% of Kusche Thinking externalising behaviours. Affective Interviews Produced sustained reduction in effectiveness scores. Strategies) • student self-reported depressive symptoms. • Coder reliability reported as.86. Table 2: (Continued) 9 Journal of Research in Special Educational Needs,  – tempo Jolivette et al. (2001) Reported fidelity data range between 98% - Author (Year) Kennedy et al. (2014) Design AT Setting Residential school Intervention Written teacher and Target Behaviour Reported outcomes • peer praise notes • Reduced inappropriate Fidelity data beha- • Fidelity data calculated viours. over 43% of sessions, Distinction between two types ranged between 50% to of written praise was not obvi- 100%. ous. • Mean of fidelity scores was more than 90%. • IOA over 93% for each student Lopata (2003) QE Self-contained school PMR (Progressive Behaviour • Muscle Relaxation) Decrease in aggression at post- • Inter-rater reliability test not present in follow-up. established during training No significant change. for identifying aggressive behaviour exceeded.85. • Reliability observation check once per classroom during pretest, post-test and follow-up periods exceeded.85. Malmgren et al. (2005) MBD General education Peer interaction Behaviour • training for paraprofessionals to • increase peer interaction in classroom Student interactions increased Interobserver agreement checked on 18% of 146 Intervention helped parapros. be total observations. more aware of their actions. • • • • • slightly. • ment on proximity data. PND: 6% student 1, 99% interobserver agree • 100% interobserver 57% student 2, agreement on rate of 33% student 3. para facilitative beha- viours. ª 2021 NASEN Table 2: (Continued) Journal of Research in Special Educational Needs,  – 10 Table 2: (Continued) ª 2021 NASEN Table 2: (Continued) Author (Year) Mason and Shriner (2008) Design Setting Multiple-probe Inclusion support classroom Intervention Self-Regulated Target Academic Reported outcomes • Strategy • Development Fidelity data All students improved persua- • 100% of lessons video sive essays. recorded and examined by Mixed results reported for post- authors. (SRSD) for test, generalisation and mainte- POW + TREE nance phases. • Fidelity of treatment: 100% for n steps checked by instructor and 98% for taperecorded steps. McDaniel et al. (2017) MBD Self-contained school Stop and Think Behaviour • • Decreased negative social beha- • Rated using 24 item cur- viours riculum checklist during Effects maintained 25% of sessions • Mean adherence was 98% • IOA of fidelity calcu lated for 2 sessions and had 100% agreement McDaniel et al. (2016) WS Self-contained school Check, Connect and Behaviour • Expect (CCE) Significant improvement on ratings. • Academic engagement improved Fidelity of teacher adher ence at 92%. • Procedural fidelity rated No significant change in aca- on 23% of check-in demic performance. activities at 92%, 22% of check-out activities at 96% and 21% of social skills activities at 98%. • IOA on 21% of weekly measures reported at 97% 11 Table 2: (Continued) Journal of Research in Special Educational Needs,  – of teacher/parapro student onset of CCE and positive beha- at the onset of CCE. • Fidelity assessed on 25% Daily Progress Reports at the vioural growth during CCE. • • Author (Year) Meyer (1999) Design AT Setting Self-contained school Intervention Function-based Target Behaviour Reported outcomes • Learning a relevant response Multiple-treatment reversal Summer Treatment Programme • All sessions were video- replacement yielded a decrease in observed taped, scored by two behaviour off-task behaviour. trained raters. • interventions Miller et al (2014) Fidelity data IOA across 35% of ses- delivered sions averaged 92% for intervention one on off-task behaviour and one in analog 96% for experimenter settings comments. Modified behaviour Behaviour • intervention where Negative behaviours lowest dur- • NR • IRR checks completed ing low-punishment condition. A = standard • No evidence supporting efficacy treatment, of high reward or combined B = low- treatment conditions. punishment treatment, C = high reward treatment and BC = combined treatment Peltier and Vannest (2018) MBD Self-contained Schema Identification Academic • Instruction • Students demonstrated improvement from baseline. on 89% of baseline and Scores decreased in maintenance 54% of intervention and but remained above baseline maintenance phases. levels. Reported as 100% across all three phases. • Overall fidelity of implementation reported as 93%. ª 2021 NASEN Table 2: (Continued) Journal of Research in Special Educational Needs,  – 12 Table 2: (Continued) ª 2021 NASEN Table 2: (Continued) Author (Year) Rafferty (2012) Design MBD Setting Summer Reading Programme Intervention Self-regulating strategy SLANT Target Academic & Reported outcomes • Behaviour and tactile All students showed gains in Fidelity data • IRR ranged from 93.8%- oral reading and increased on- 97.9% for observation of task behaviour. on-task behaviour. • prompting device Checklists used for treatment integrity and accuracy of student self- during two monitoring observations range: 71.4%-100%. • Teacher instructional behaviour rated at 100% across all sessions. Scruggs and Osguthorpe (1986) QE classrooms Cross-age Peer Academic • Tutoring Academic gains reported for • NR • NR tutors and tutees. Attitudinal gains observed for tutees. Scruggs and Osguthorpe (1986) Study 2 QE Special education Peer Tutoring Academic • classrooms Academic gains for tutors and tutees. • No significant changes in Attitude measure Table 2: (Continued) 13 Journal of Research in Special Educational Needs,  – Study 1 Special education Author (Year) Trussel et al. (2008) Design MBD Setting Self-contained school Intervention Function-based Target Behaviour Reported outcomes • Problem behaviours Fidelity data were • reduced across all three subjects replacement Weekly observation and Likert scale used to moni- behaviour tor interventions and IOA data collected during treatment targeted classroom 70% interventions analysis of mean the integrity. functional observations, agreement: 85% across all subjects. • 56% of sessions included a second baseline, observer for classroom and FBA interventions. • Mean agreement across all data points, 96%. Weeden et al. (2016) Reversal Self-contained Class-wide Function- Behaviour Related • • Intervention Team Increased on-task behaviour, increased levels of • teacher praise and • Procedural fidelity data collected for 75% of sessions. decreased levels of teacher rep- • rimands. Independent variable implemented with average of 96% fidelity. Wehby et al. (2003) MBD Self-contained school Open Court Reading curriculum and Peer-Assisted Academic & • Behaviour • Learning Strategies • Moderate gains in reading • 21% of first grade and achievement for both groups. 33% of kindergarten OCR No increases in standardised sessions were observed for scores implementation fidelity. No decreases in problem beha- • viours were observed. Overall MN fidelity of treatment 97%. • 67% of PALS sessions were observed • Overall mean implementa- ª 2021 NASEN tion fidelity 96%. Table 2: (Continued) Journal of Research in Special Educational Needs,  – 14 Table 2: (Continued) ª 2021 NASEN Table 2: (Continued) Author (Year) Wehby et al. (2005) Design MBD Setting Self-contained Intervention Scott Foresman Target Academic Reported outcomes • Moderate but inconsistent gains Fidelity data • Treatment integrity data were shown in nonsense word was collected on 69% of and Phonological fluency, letter naming fluency the sessions in group 1 Awareness and onset fluency. and 28% of the sessions in reading programme group 2. Training for Reading (PATR) • Mean session integrity ratings: Trainers: 99.58% group 1 and 100% group 2. External observers 99.45% group 1 and 100% group 2. 15 Journal of Research in Special Educational Needs,  – Notes: MBD = multiple baseline design; QE = quasi-experimental; RCT = randomised control design; NR = not reported; EBD = Emotional and Behavioural Disorder; LD = Learning Disability; IOA = Interobserver agreement; IRR = Inter-rater Reliability. Journal of Research in Special Educational Needs,  – social skills (Blair et al., 2000), progressive muscle relaxation (Lopata, 2003), paraprofessional training in methods to increase peer interactions (Malmgren et al., 2005), a function-based replacement behaviour intervention (Meyer, 1999), a function-based replacement intervention paired with targeted classroom interventions (Trussel et al., 2008) and class-wide function-related intervention team (Weeden et al., 2016). Three of the eleven studies (27.2%) focused on academic interventions, all of which included some element of explicit instruction: a mathematics problem-solving intervention (Alter et al., 2011), a comparison of different methods for providing a reading model (Dawson et al., 2000) and the kindergarten version of peer-assisted learning strategies (KPALS; Wehby et al., 2003). One study (Wehby et al., 2005) investigated intervention effects on academic performance and behaviour. In this study, students were provided a multi-component reading intervention (e.g. Open Court Reading, PALS). Design characteristics. Twenty-three studies (76.6%) employed a single case design. Multiple baseline designs were most commonly used (n = 10, 33.3%), followed by alternating treatment (n = 6, 20%), hybrid (n = 3, 10%), reversal (n = 3, 10%) and multiple probe (n = 1, 3.3%) designs. The remaining studies (n = 7) used some variation of a group design (20%). Four studies (13.3%) employed a quasi-experimental design, two studies (6.6%) employed a randomised control group design, and one study (3.3%) used a within-subject design. RQ2: WWC coding All intervention studies (n = 30) were screened to determine whether they employed a design eligible for WWC coding (e.g. comparison group and single case). Two studies (6.6%; McDaniel et al., 2016; Miller et al., 2014) employed ineligible designs (interrupted time series and within-subjects repeated measures, respectively). Four studies (13.3%; Benner et al., 2012; Kam et al., 2004; Scruggs and Osguthorpe, 1986 Study 1 & 2) used an eligible design but did not disaggregate outcomes for students with EBD. One study used an eligible design and disaggregated outcomes for students with EBD but did not do so for children with EBD in EC to grade 2 (Weeden et al., 2016). As a result, we were unable to evaluate these studies according to the quality indicators. In total, twenty-three intervention studies (76.6%; 2 comparison group, 21 single case) were evaluated according to the applicable WWC design standards. Table 3 reports WWC quality indicator coding for comparison group studies. Table 4 reports WWC quality indicator coding for single case design studies. Ten (30%) of 30 investigations met WWC design standards with or without reservations. Five studies (16.6%) met design standards without reservations (Dawson et al., 2000; Dwyer et al., 2012; Kennedy et al., 2014; Lopata, 2003; Peltier and Vannest, 2018). When analysing Peltier and 16 Vannest (2018), we considered the different schema types assessed after baseline as one intervention phase. Five studies (16.6%) met design standards with reservations (Alter et al., 2011; Dunlap et al., 1996; Hagan-Burke et al., 2015, study 1; Meyer, 1999; Trussel et al., 2008). Thirteen studies (43.3%) did not meet design standards (Adkins and Gavins 2012; Beck et al., 2009; Blair et al., 2000; Clair et al., 2018; Cochran et al., 1993; Falk and Wehby, 2001; Jolivette et al., 2001; Malmgren et al. 2005; Mason and Shriner, 2008; McDaniel et al., 2017; Rafferty, 2012; Wehby et al., 2003; Wehby et al., 2005). Single case studies were commonly rated as not meeting standards due to employing a design that did not provide an opportunity to demonstrate an intervention effect at three different points in time. Studies meeting WWC design standards Dawson et al. (2000) compared the effects of a computerbased model and a teacher provided model to a no modelling condition. This study was conducted in a public school resource room setting and included four students with EBD, two in first grade and two in second grade. All students also received instruction in general education classrooms. In the no modelling condition, students read a passage independently. In the computer model condition, students first followed along in the passage as it was read by the computer and then read the passage independently. In the teacher modelling condition, students first followed along as the teacher read the passage and then read it independently. Intervention sessions were conducted in a one-to-one setting. When not receiving oneto-one instruction, students were given the option of working on a computer with headphones on or reading a self-selected book. In this investigation, the teacher model and computer model were both superior to the no modelling condition for mean rate of words read correctly per minute. However, students read more words correctly on average in the teacher model condition. Dwyer et al. (2012) investigated the effects of a functionbased replacement behaviour intervention with three students with ED ranging in age from 7 to 8 years of age. Students received instruction in a self-contained class within an urban elementary school. A functional behavioural assessment consisting of a record review, teacher interview and direct observation was completed for each student by the researchers. For all three students, problem behaviours were prevalent during independent seatwork and whole group instruction, with challenging behaviours performed to escape these activities. In this study, students were provided explicit instruction in appropriate methods for asking for assistance and asking for a break. During intervention sessions, students were required to ask for help if needed, ask for a break if needed, ask for either assistance or break if needed, or not ask to use a replacement behaviour, depending on the session type that was selected by researchers. During sessions in which a replacement behaviour was expected, students ª 2021 NASEN Journal of Research in Special Educational Needs,  – Table 3: WWC ratings for comparison group studies Cochran et al. (1993) Lopata (2003) A primary analysis of the effect of an intervention Yes Yes Eligible design QED Cluster RCT Random process? No Yes If yes go to sample attrition. If no, go to baseline equivalence. Sample attrition Is the combination of overall and differential attrition high? No Baseline equivalence No Determination Does not meet Meets Standards Notes: WWC = What Works Clearinghouse; RCT = Randomised Controlled Trial; QED = Quasi-experimental Design. were told what replacement behaviour was acceptable and given a visual cue card as a reminder prior to each intervention session. In this study, each student experienced a training period until they were able to perform the required replacement behaviour for five consecutive opportunities over the course of two consecutive days. Findings suggest that the replacement behaviour interventions were effective at decreasing off-task behaviour. Kennedy, Jolivette and Ramsey (2014) investigated the effects of teacher praise notes and peer praise notes in a residential school for students with EBD. Participants were eight students with EBD in grades two to four, all of whom met eligibility criteria for ED. In this study, each student received a functional behavioural assessment to confirm that their challenging behaviours were maintained by attention. Functional behavioural assessments, performed by the researchers, consisted of a record review, teacher interviews and direct observation. Intervention sessions were conducted during art class. During the teacher praise note condition, the teacher wrote a behaviour-specific praise statement on a star-shaped piece of paper for each student. Students received the praise notes before the last scheduled art activity, which occurred when 15 minutes remained in the class. During the peer praise note condition, students were given a blank star-shaped piece of paper when 15 minutes remained in the class and directed to write a praise statement for any peer in the class. Peer praise statements were reviewed for appropriate content prior to being given to students and each student received at least one peer praise note each session. In this study, both types of praise notes were effective at decreasing challenging behaviour. Lopata (2003) investigated the effects of a progressive muscle relaxation intervention on the aggressive behaviour of 24 children with EBD who attended a day school treatment programme. Students were six to nine years of age. Classroom teachers were trained in the intervention protocol by researchers, which was based on procedures by Jacobson (1942, as cited in Lopata, 2003) and ª 2021 NASEN provided in 40 minute sessions five days per week for four weeks. Children receiving intervention were guided through a series of prompts to tense up, maintain the tension and then release the tension for specific muscles and muscle groups. The researchers reported the intervention as effective at decreasing aggression. Although intervention effects were not maintained at follow-up, aggressive incidents were below the levels reported prior to intervention. Peltier and Vannest (2018) investigated the effects of schema based instruction on the word problem-solving skills of four second grade students with EBD who were served in a self-contained classroom. In this investigation, a researcher served as the interventionist, with schema instruction in the STAR method delivered to students in a one-to-one instructional setting. Instruction followed a modelling, guided practice and independent practice instructional sequence, with students provided additional guided practice if they required more than one prompt. During independent practice, students had the option for the researcher read the word problem aloud to them. Students were provided corrective feedback and additional independent practice when errors were made. A token economy was also employed and students earned a check mark for each question attempted and a tangible item for attempting all assigned questions. The researchers reported that the intervention was effective, although results and maintenance effects were variable across students. Studies meeting WWC design standards with reservations Alter et al. (2011) investigated the effects of strategy instruction on the mathematics performance of three students who attended a self-contained class for students with EBD. One student was in first grade and two students were in fourth grade. The classroom teacher was provided four hours of training over the course of two sessions. Training included instruction in Making a Table or Chart, Guess and Check, and Draw a Picture strategies and following a scripted lesson protocol. A one-hour follow-up training focusing on cueing methods 17 Adkins and Gavins (2012) The independent variable is systematically manipulated Yes Alter et al. (2011) Yes Beck et al. (2009) Yes Blair et al. (2000) Yes Clair et al. (2018) Yes Dawson et al. (2000) Yes Dunlap et al. (1996) Yes Dwyer et al. (2012) Yes Outcome variables measured systematically Yes Yes Yes Yes Yes Yes Yes Yes Inter-assessor agreement on at least 20% of data points in Yes Yes Yes Yes No Yes Yes Yes Yes Yes Yes* Yes Yes Yes Yes Yes No Yes Yes* No Yes Yes Yes Yes Minimum number of data points No Yes Yes* No Yes Yes Yes Yes Determination Does not meet With Does not meet Does not meet Does not meet standards standards standards each condition Inter-assessor agreement must be .80 for percentage agreement and .60 for kappa At least three attempts to demonstrate an intervention effect reservations standards Meets With standards reservations Meets standards*** Falk and Wehby (2001) Hagan-Burke et al. (2015) Study 1 Jolivette et al. (2001) Kennedy et al. (2014) Malmgren et al. (2005) Mason and Shriner (2008) McDaniel et al. (2017) The independent variable is systematically manipulated Yes Yes Yes Yes Yes Yes Yes Outcome variables measured systematically No Yes Yes Yes Yes Yes Yes Inter-assessor agreement on at least 20% of data points in No Yes Yes Yes No Yes Yes No Yes Yes Yes Yes Yes Yes each condition Inter-assessor agreement must be .80 for percentage agreement and .60 for kappa Table 4: (Continued) Journal of Research in Special Educational Needs,  – 18 Table 4: WWC ratings for single case design studies ª 2021 NASEN ª 2021 NASEN Table 4: (Continued) Falk and Wehby (2001) At least three attempts to demonstrate an intervention Yes Hagan-Burke et al. (2015) Study 1 Jolivette et al. (2001) Yes No Kennedy et al. (2014) Malmgren et al. (2005) Mason and Shriner (2008) McDaniel et al. (2017) Yes Yes Yes No effect Minimum number of data points No Yes Yes Yes Yes No** Yes Determination Does not meet With reservations Does not meet Meets standards Does not meet Does not meet Does not meet standards standards standards standards standards Meyer (1999) The independent variable is systematically manipulated Yes Peltier and Vannest (2018) Yes Rafferty (2012) Yes Trussel et al. (2008) Yes Wehby et al. (2003) Wehby et al. (2005) Yes Yes Outcome variables measured systematically Yes Yes Yes Yes Yes Yes Inter-assessor agreement on at least 20% of data points in each Yes Yes Yes Yes No Yes Yes Yes Yes Yes Yes Yes At least three attempts to demonstrate an intervention effect Yes Yes No No No No Minimum number of data points Yes Yes Yes Yes Yes Yes Determination With Meets standards Does not meet With reservations Does not meet Does not meet standards standards condition Inter-assessor agreement must be .80 for percentage agreement and .60 reservations standards Notes: WWC = What Works Clearinghouse. *Met criteria for one participant. **Did not meet additional criteria for multiple probe designs. ***Evaluated as an alternating treatment design. 19 Journal of Research in Special Educational Needs,  – for kappa Journal of Research in Special Educational Needs,  – was also provided so that the intervention better addressed teacher and student needs. Intervention sessions were 30 minutes in duration, which consisted of 15 minutes of teacher instruction and 15 minutes of student practice. In this investigation, the researchers reported that the intervention was effective at improving time on task and correct problem completion. Dunlap et al. (1996) investigated the effects of functionbased interventions on the problem behaviour and task engagement of three students with SED. Participants attended specialised classrooms for EBD and were in first grade, second grade and fourth grade. Functional behavioural assessments, completed by the researchers, consisted of a record review, teacher interviews, student interview and parent/guardian interview, and classroom observations. Assessments were used to develop multicomponent interventions that focused on modifying class assignments. Modifications included use of enlarged print, shortened assignments, incorporating student choice, use of visuals and underlining and highlighting of words on worksheets. Students were also reminded to ask for help if they needed it and to inform teachers when they completed their assigned tasks. In this investigation, the researchers reported positive effects on task engagement and instances of challenging behaviour. Hagan-Burke et al. (2015) investigated the effects of slowing the pace of academic instruction on the task engagement of a first grader with BD and ADHD. Functional behaviour assessments were performed by the researchers and included record reviews, teacher interviews and classroom observation. The intervention was provided by the special education teacher in a special education classroom to a small group of students. In this investigation, the pace at which students were prompted to chorally respond during circle time was adjusted from the pace of instruction that was observed in the general education classroom to a pace that was thought to be more ideal for the target student. In this investigation, the researchers reported positive effects. However, it should be noted that the general education teacher did not consent to participate in this investigation, resulting in the intervention being provided in the special education classroom. Meyer (1999) investigated the effects of function-based interventions with four students who attended a school for children with learning and behavioural disabilities. Two participants were in first grade and two were in fourth grade. In this study, the researchers altered the amount of attention students received as well as the task difficulty to determine the conditions that most likely predicted high rates of off-task behaviour. Students were also taught a method for recruiting adult attention and a method for asking for assistance. In addition, researchers implemented the functional analysis for each student in a separate room. Meyer reported that functional analysis 20 may be an effective method for designing interventions for this student population. Trussel et al. (2008) investigated the effects of a function-based intervention and a targeted behavioural intervention on the problem behaviour of three students who attended an alternative public school for students with significant behavioural and mental health needs. Participants were in first, third and fifth grade. Researchers completed the functional behavioural assessments, which included classroom observations, teacher interviews and a functional analysis. Targeted classroom interventions varied by student, as they attended different classes. Targeted intervention included various methods such as posting student work, teaching replacement behaviours and reinforcing student hand raising. Function-based intervention also varied by student and included explicit instruction in replacement behaviours (e.g. ask for time to work with peers and adults, request opportunity to model a task for peers), increased opportunities to respond, increased use of positive reinforcement and providing opportunities for student choice. Replacement behaviour instruction was performed using the Skill Streaming (McGinnis and Goldstein, 1997) curriculum. In this investigation, researchers reported positive effects on student behaviour. Discussion Systematic reviews are commonly performed to report existing research, identify effective practices and make recommendations for future studies that are necessary to inform the field of special education. It is imperative that research-based practices are identified for young children with EBD in order to improve the negative trajectories that they experience (Bulotsky-Shearer et al., 2010; Conroy et al., 2015). However, researchers have not previously systematically reviewed and evaluated studies investigating the effects of interventions on young children with EBD. This is a notable gap in the research because early intervention for students with EBD is best practice (Garwood et al., 2017b). The purpose of this investigation was to describe the peer-reviewed research investigating the effects of interventions on young children with EBD. We also sought to identify promising practices for improving child outcomes by critically evaluating identified studies according to the WWC design standards and making recommendations for practice based solely on those studies the met design standards with or without reservations. RQ1: Study characteristics Considering the broad nature of our article selection criteria (e.g. intervention studies with at least one young child formally identified with EBD), it was surprising to find so few intervention studies available in the literature. The lack of attention to children who are six years of age or younger is particularly problematic, considering early intervention is the key to positive outcomes for children ª 2021 NASEN Journal of Research in Special Educational Needs,  – with behaviour problems (Walker et al., 2004). However, the limited number of studies with young children with EBD who were between the ages of EC and six is likely an artefact of the challenges associated with identifying young children (see Conroy and Brown, 2004; Wakschlag et al., 2007). It is possible that the current push for universal pre-K and improved access to high quality health care in the United States could eventually result in an increase in the number of children who are formally identified with EBD. Improving child and family access to professionals (e.g. educational and medical) with expertise in the characteristics of EBD may result in more accurate and timely identification. The majority of participants within this pool of studies were male (87.2% of participants for whom this information reported). This finding is not surprising as females are more likely to display internalising behaviours than externalising, which are more readily identified (Garwood et al., 2017a; Poulou, 2015). Prior research has shown boys are more often the target of teachers’ attention, for both positive and negative reasons (Beaman et al., 2006; Rice and Yen, 2010). Therefore, the predominance of externalising behaviours on the part of boys could explain the gender differences in study samples. Teachers, who are often called upon to identify students with EBD through the use of behavioural screening instruments, tend to focus much more on the externalising behaviours that disrupt the classroom and are far more common in boys than girls (Harrison et al., 2012). The majority of studies meeting article selection criteria (n = 23; 76.6%) were conducted in specialised school settings. This finding is concerning, as the majority of students with EBD are served in general education or inclusive classrooms. Researchers have previously noted the dearth of intervention studies that are applicable to students with EBD who are educated in general education settings (McKenna et al., 2019; Scott and Alter, 2017). This finding may reflect the research that suggests inclusion of students with EBD may have negative effects on the school performance of students who do not have a disability (Farmer and Hollowell, 1994; Fletcher, 2010; Gottfried, Egalite, and Kirksey, 2016). Thus, educating young children with EBD in inclusive classrooms may pose a significant challenge for practitioners due to the absence of intervention research, the potential negative effects of EBD and the likely need to high levels of expertise and resources to promote positive outcomes. RQ2: Quality indicator coding Of those studies meeting design standards with or without reservations, five (Dunlap et al., 1996; Dwyer et al., 2012; Hagan-Burke et al., 2015 study 1; Meyer, 1999; Trussel et al., 2008) investigated the effects of functionbased interventions. Function-based interventions tended to involve curricular and instructional adaptations, reinforcement and student instruction in replacement ª 2021 NASEN behaviours. Researchers in these investigations also reported positive effects on student behaviour. Researchers have previously noted that function-based interventions have positive effects on the behaviour of students with EBD (Gage et al., 2012). Function-based interventions that include replacement behaviour training may also an effective practice for improving student behaviour (McKenna et al., 2015; McKenna et al., 2017). Function-based replacement behaviour interventions involve conducting a functional behavioural assessment to identify behaviours or skills that are functionally equivalent (e.g. meet the same need) or incompatible to the performance of challenging behaviours in addition to the contextual factors that predict challenging behaviour. In addition to changes to the environment, students are provided direct instruction on how and when to perform the identified replacement behaviours. Students also typically receive some form of reinforcement for their performance. However, in the present review, all function-based replacement behaviour intervention studies with designs that permitted some degree of causal inference were conducted in substantially separate settings. As a result, we are unable to estimate the effectiveness of function-based replacement interventions for young children with EBD who are educated in inclusive classrooms. This finding is also consistent with previous research, which suggested that function-based interventions are not an evidencebased practice for use in general education settings (Scott and Alter, 2017). Implications for practice Practitioners must use research evidence in concert with their professional values, judgement and training when making decisions regarding how best to serve their students (e.g. evidence-based practice as a process; Cook et al., 2016b; Cook et al., 2016a). Findings from this systematic review suggest that there is limited research to inform practice for young children who are formally identified with EBD. Nevertheless, practitioners must make decisions regarding how best to serve students in the absence of research. Findings from this review suggest that function-based interventions may be an effective practice for improving student behaviour. However, function-based interventions included in this review tended to involve a high degree of researcher support and implementation. As result, EC and school-based practitioners may be in need of additional resources, supports and expertise to perform functional behavioural assessments and implement function-based interventions with fidelity. This need for additional technical assistance and support may be particularly salient to their use and effectiveness in inclusive classrooms (see McKenna et al., 2017). Furthermore, the use of function-based interventions to support young children with EBD would place a greater expectation on schools than what is required by special education law (see Yell, 2019). At present, schools in the United States are only required to ‘consider’ the use of 21 Journal of Research in Special Educational Needs,  – positive behaviour supports such as function-based interventions. A reauthorisation of federal special education law may be necessary to ensure schools have the necessary resources to for this recommendation to be feasible in at least some educational settings (see McKenna and Brigham, 2021). Progressive muscle relaxation may also be an effective practice for promoting self-regulation (Lopata, 2003). In fact, previous research investigating the effects of functionbased interventions included training in progressive muscle relaxation as part of replacement behaviour training (McKenna et al., 2017). Thus, it may be effective to pair function-based interventions with progressive muscle relaxation for those young children with EBD who have difficulties with self-regulation and with externalising challenging behaviours. Again, investigations meeting selection criteria tended to involve a high level of researcher support. As a result, school-based practitioners may need additional support in the form of technical assistance and resources for interventions to be implemented with sufficient quality, consistency and intensity in typical settings. This proves particularly true for interventions that are informed by functional behavioural assessments, which were performed by researchers across studies. Limitations Two limitations are associated with this study. First, we did not include grey literature in our search procedures. Researchers have previously expressed concern with the omission of grey literature from systematic reviews (see Gage et al., 2017). However, we sought to base our findings exclusively on studies that were evaluated by peer review. Peer review is considered a marker for research quality. Second, we coded studies according to a rigorous set of quality indicators. Findings from this systematic review may have differed if we evaluated studies according to another set of expert panel recommendations (e.g. Council for Exceptional Children, 2014). Future research Findings from this systematic review point towards six recommendations for future investigations. First, researchers should continue the difficult and much needed work associated with the timely and accurate identification of young children with EBD. Identification is a pre-requisite not only to participant inclusion in intervention studies, but to the timely and effective delivery of research-based supports in practice. Second, additional studies investigating the effects of interventions for young children with EBD are needed as few studies in this review met WWC design standards with or without reservations. Although randomised control trials continue to be the gold standard in intervention research, the aforementioned issues with child identification likely preclude the use of this methodology. To address this potential issue, we suggest that researchers employ single case designs, which has a robust history across fields including special education due to their ability to perform 22 rigorous evaluations with small samples of participants (see Barlow and Hersen, 1973; Horner et al., 2005; Radley et al., 2020). When conducting single case investigations, we suggest that researchers employ designs informed by expert panel recommendations so that some degree of causal inference can be made from the introduction of independent variables and observed changes in student performance. This need to apply quality indicators is particularly salient to research focusing on academics such as early reading skills. Students with disabilities are entitled to a free appropriate public education (FAPE), and this is determined by the degree to which special education services confer appropriate benefit rather than the location (e.g. student placement) in which services are provided. In instances in which a child’s disability adversely affects academic performance, special education services must confer appropriate benefit in these areas. Third, future investigations should report information on student ethnicity and socioeconomic status, as this information was commonly not reported. This information is necessary so that researchers and practitioners alike can make informed decisions regarding the degree to which studies are applicable to their school partners and students. Fourth, researchers are encouraged to disaggregate study results for young children with EBD so that researchers can determine the degree to which this student population benefited from interventions. Researchers can also publish their data online as supplementary materials so that secondary or descriptive analyses can be performed. Fifth, replication studies are needed to develop a more robust research base for the use of practices included in studies that reported positive effects. Study replication is essential to the identification of evidence-based practices (Travers et al., 2016). Specifically, replication studies are needed to explore the potential effectiveness of function-based interventions, progressive muscle relaxation and schema based instruction for young children with EBD. Lastly, researchers are encouraged to continue to develop lines of research that inform the responsible inclusion of students with EBD. Inclusive instruction for students with disabilities is a complex endeavour, with varying degrees to which practice resembles actual inclusion rather than merely integration (see Brigham et al., 2016; Solis et al., 2012). The identification and dissemination of research-based practices is essential to better ensure that practice is driven by science rather than philosophy and/or expediency. *Address for correspondence John William McKenna, University of Massachusetts Lowell, 850 Broadway St., Lowell, MA 01854, USA. Email: john_mckenna@uml.edu. ª 2021 NASEN Journal of Research in Special Educational Needs,  – References References marked with an asterisk indicate studies included in the literature review. *Adkins, M. H. & Gavins, M. V. (2012) ‘Self-Regulated strategy development and generalization instruction: Effects on story writing and personal narratives among students with severe emotional and behavioral disorders.’ Exceptionality, 20, pp. 235–249. https:// doi.org/10.1080/09362835.2012.724625. *Alter, P., Brown, E. T. & Pyle, J. (2011) ‘A strategyy based intervention to improve math word problemsolving skills of students with emotional and behavioral disorders.’ Education and Treatment of Children, 34, pp. 535–550. https://doi.org/10.1353/etc. 2011.0028. Barlow, D. & Hersen, M. (1973) ‘Single-case experimental designs: Use in applied clinical research.’ Archives of General Psychiatry, 29, pp. 319–325. https://doi.org/10.1001/archpsyc.1973. 04200030017003. Beaman, R., Wheldall, K. & Kemp, C. (2006) ‘Differential teacher attention to boys and girls in the classroom.’ Educational Review, 58, pp. 339–366. https://doi.org/10.1080/00131910600748406. *Beck, M., Burns, M. & Lau, M. (2009) ‘The effect of preteaching reading skills on the on-task behavior of children identified with behavior disorders.’ Behavioral Disorders, 34, pp. 91–99. https://doi.org/ 10.1177/019874290903400203. *Benner, G. J., Nelson, J. R., Sanders, E. A. & Ralston, N. C. (2012) ‘Behavior intervention for students with externalizing behavior problems: Primary level standard protocol.’ Exceptional Children, 78, pp. 181– 198. https://doi.org/10.1177/001440291207800203. *Blair, K. C., Umbreit, J. & Eck, S. (2000) ‘Analysis of multiple variables related to young child’s aggressive behavior.’ Journal of Positive Behavior Interventions, 2, pp. 33–39. https://doi.org/10.1177/ 109830070000200105. Brigham, F., Ahn, S., Stride, A. & McKenna, J. (2016) ‘FAPE accompli: Misapplication of the principles of inclusion and students with EBD.’ In J. P. Bakken & F. E. Obiakor (eds) Advances in Special Education: General and Special Education Inclusion in an Age of Change: Impact on Students with Disabilities, Vol. 32, pp. 31–47.Bingley, UK: Emerald Group Publishing Limited. Bulotsky-Shearer, R. J., Dominguez, X., Bell, E. R., Rouse, H. L. & Fantuzzo, J. W. (2010) ‘Relations between behavior problems in classroom social learning and learning situations and peer social competence in Head Start and kindergarten.’ Journal of Emotional and Behavioral Disorders, 18, pp. 195– 210. https://doi.org/10.1177/1063426609351172. Carter, A., Wagmiller, R., Gray, S., McCarthy, K., Horwitz, S. & Briggs-Gowan, M. (2010) ‘Prevalence of DSM-IV disorder in representative, healthy birth ª 2021 NASEN cohort at school entry: Sociodemographic risks and social adaptation.’ Journal of the American Academy of Child & Adolescent Psychiatry, 49, pp. 686–689. https://doi.org/10.1016/j.jaac.2010.03.018. *Clair, E., Bahr, M., Quach, H. & LeDuc, J. (2018) ‘The Positive Plus Program: Affirmative classroom management to improve student behavior.’ Behavioral Interventions, 33, pp. 221–236. https://doi.org/10. 1002/bin.1632. *Cochran, L., Feng, H., Cartledge, G. & Hamilton, S. (1993) ‘The effects of cross-age tutoring on the academic achievement, social behaviors, and selfperceptions of low-achieving African American males with behavioral disorders.’ Behavioral Disorders, 18, pp. 292–302. https://doi.org/10.1177/ 019874299301800402. Conroy, M. & Brown, M. (2004) ‘Early identification, prevention, and early intervention with young children at risk for emotional or behavioral disorders: Issues, trends, and a call for action.’ Behavioral Disorders, 29, pp. 224–236. https://doi.org/10.1177/ 019874290402900303. Conroy, M., Sutherland, K., Algina, J., Wilson, R., Martinez, J. & Whalon, K. (2015) ‘Measuring teacher implementation of the BEST in CLASS Intervention Program and corollary child outcomes.’ Journal of Emotional and Behavioral Disorders, 23, pp. 144– 155. https://doi.org/10.1177/1063426614532949. Cook, B., Collins, L., Cook, S. & Cook, L. (2016b) ‘A replication by any other name: A systematic review of replicative intervention studies.’ Remedial and Special Education, 37 (4), pp. 223–234. https://doi.org/10. 1177/0741932516637198. Cook, B., Cook, S. & Collins, L. (2016a) ‘Terminology and evidence-based practice for students with emotional and behavioral disorders: Exploring some devilish details.’ Beyond Behavior, 25 (2), pp. 4–13. https://doi.org/10.1177/107429561602500202. Costello, E., Egger, H. & Angold, A. (2005) ‘Ten-year research update review: The epidemiology of child and adolescent psychiatric disorders: I. Methods and public health burden.’ Journal of the American Academy of Child and Adolescent Psychiatry, 44, pp. 972–986. https://doi.org/10.1097/01.chi.0000172552.41596.6f. Council for Exceptional Children (2014) Council for Exceptional Children Standards for Evidence-Based Practices in Special Education. Arlington, VA: Council for Exceptional Children. *Dawson, L., Ven, M. L. & Gunter, P. L. (2000) ‘The effects of teachers versus computer reading models.’ Behavioral Disorders, 25, pp. 105–1013. https://doi. org/10.1177/019874290002500202. *Dunlap, G., White, R., Vera, A., Wilson, D. & Panacek, L. (1996) ‘The effects of multi-component, assessment-based curricular modifications on the classroom behavior pf children with emotional and behavioral disorders.’ Journal of Behavioral 23 Journal of Research in Special Educational Needs,  – Education, 6, pp. 481–500. https://doi.org/10.1007/bf 02110518. Dwyer, K., Rozewski, D. & Simonsen, B. (2012) ‘A comparison of function-based replacement behaviors for escape-motivated students.’ Journal of Emotional and Behavioral Disorders, 20 (2), pp. 115–125. https://doi.org/10.1177/1063426610387432. Every Student Succeeds Act (ESSA) (2015) Pub. L. 114– 95, § 1177. *Falk, K. B. & Wehby, J. H. (2001) ‘The effects of peerassisted reading skills of young children with emotional or behavioral disorders.’ Behavioral Disorders, 26, pp. 344–359. https://doi.org/10.1177/ 019874290102600404. Farmer, T. W. (2013) ‘When universal approaches and prevention services are not enough: The importance of understanding the stigmatization of special education for students with EBD. A response to Kauffman and Badar.’ Behavioral Disorders, 39, pp. 32–42. https:// doi.org/10.1177/019874291303900105. Farmer, T. W. & Hollowell, J. H. (1994) ‘Social networks in mainstream classrooms: Social affiliations and behavioral characteristics of students with EBD.’ Journal of Emotional and Behavioral Disorders, 2 (3), pp. 143–155. Fletcher, J. (2010) ‘Spillover effects of inclusion of classmates with emotional problems on test scores in early elementary school.’ Journal of Policy Analysis and Management, 29, pp. 69–83. https://doi.org/10. 2139/ssrn.1424191. Forness, S. R., Freeman, S. F. N., Paparella, T., Kauffman, J. M. & Walker, H. M. (2012a) ‘Special education implications of point and cumulative prevalence for children with emotional or behavioral disorders.’ Journal of Emotional and Behavioral Disorders, 20, pp. 4–18. https://doi.org/10.1177/ 1063426611401624. Forness, S. R., Kim, J. & Walker, H. M. (2012b) ‘Prevalence of students with EBD: Impact on general education.’ Beyond Behavior, 21, pp. 3–10. Gage, N., Cook, B. & Reichow, B. (2017) ‘Publication bias in special education meta-analysis.’ Exceptional Children, 83 (4), pp. 428–445. Gage, N., Lewis, T. & Stichter, J. (2012) ‘Functional behavioral assessment-based interventions for students with or at risk for emotional and/or behavioral disorders in school: A hierarchical linear modeling meta-analysis.’ Behavioral Disorders, 37 (2), pp. 55– 77. https://doi.org/10.1177/019874291203700201. Garwood, J. D., McKenna, J. W. & Ciullo, S. (2020) ‘Reading instruction with embedded behavioral supports for children with emotional and behavioral disorders.’ Beyond Behavior, 29 (1), pp. 6–17. https:// doi.org/10.1177/1074295619900380. Garwood, J., McKenna, J., Roberts, G., Ciullo, S. & Shin, M. (2019) ‘Social studies content knowledge interventions for students with emotional and behavioral disorders: A meta-analysis.’ Behavior 24 Modification, 45 (1), pp. 147–176. https://doi.org/10. 1177/0145445519834622 Garwood, J. D., Varghese, C. & Vernon-Feagans, L. (2017a) ‘Internalizing behaviors and hyperactivity/ inattention: Consequences for young struggling readers, and especially boys.’ Journal of Early Intervention, 39, pp. 218–235. https://doi.org/10.1177/ 1053815117706524. Garwood, J. D., Vernon-Feagans, L., & the Family Life Project Key Investigators (2017b) ‘Classroom management affects literacy development of students with emotional and behavioral disorders.’ Exceptional Children, 83, pp. 123–142. https://doi.org/10.1177/ 0014402916651846. Gottfried, M. A., Egalite, A. & Kirksey, J. (2016) ‘Does the presence of a classmate with emotional/behavioral disabilities link to other students’ absences in kindergarten?’ Early Childhood Research Quarterly, 36, pp. 506–520. https://doi.org/10.1016/j.ecresq.2016. 02.002. *Hagan-Burke, S., Gilmour, M. W., Gerow, S. & Crowder, W. C. (2015) ‘Identifying academic demands that occasion problem behaviors for students with behavior disorders: Illustrations at the elementary school level.’ Behavior Modification, 39, pp. 215–241. https://doi.org/10.1177/ 0145445514566505. Harrison, J. R., Vannest, K., Davis, J. & Reynolds, C. (2012) ‘Common problem behaviors of children and adolescents in general education classrooms in the United States.’ Journal of Emotional and Behavioral Disorders, 20, pp. 55–64. https://doi.org/10.1177/ 1063426611421157. Hoover, S. D., Kubicek, L. F., Rosenberg, C. R., Zundel, C. & Rosenberg, S. A. (2012) ‘Influence of behavioral concerns and early childhood expulsions on the development of early childhood mental health consultation in Colorado.’ Infant Mental Health Journal, 33, pp. 246–255. https://doi.org/10.1002/ imhj.21334. Horner, R. H., Carr, E. G., Halle, J., Mcgee, G., Odom, S. & Wolery, M. (2005) ‘The use of single-subject research to identify evidence-based practice in special education.’ Exceptional Children, 71 (2), pp. 165– 179. https://doi.org/10.1177/001440290507100203. Individuals with Disabilities Education Improvement Act (2004) 20 U.S.C § 1400. *Jolivette, K., Wehby, J. H., Canale, J. & Massey, G. (2001) ‘Effects of choice-making opportunities on the behavior of students with emotional and behavioral disorders.’ Behavioral Disorders, 26, pp. 131–145. https://doi.org/10.1177/019874290102600203. *Kam, C., Greenberg, M. T. & Kusche, C. A. (2004) ‘Sustained effects of the PATHS curriculum on the social and psychological adjustment of children in special education.’ Journal of Emotional and Behavioral Disorders, 12, pp. 66–78. https://doi.org/ 10.1177/10634266040120020101. ª 2021 NASEN Journal of Research in Special Educational Needs,  – Kauffman, J. & Landrum, T. (2013) Characteristics of Emotional and Behavioral Disorders of Children and Youth. Boston: Pearson/Merrill. *Kennedy, C., Jolivette, K. & Ramsey, M. (2014) ‘The effects of written teacher and peer praise notes on the inappropriate behaviors of elementary students with emotional and behavioral disorders in a residential school.’ Residential Treatment for Children & Youth, 31, pp. 17–40. https://doi.org/10.1080/0886571X. 2014.878577. Kratochwill, T., Hitchcock, J., Horner, R., Levin, J., Odom, S., Rindskoph, D. & Shadish, W. (2010) ‘Single-case designs technical documentation.’ What Works Clearinghouse website <http://ies.gov/ncee/ wwc/pdf/wwc_scd.pdf>. Kulkarni, T. & Sullivan, A. (2019) ‘Relationship between behavior at school entry and services received in third grade.’ Psychology in the Schools, 56 (5), pp. 809– 823. https://doi.org/10.1002/pits.22231. *Lopata, C. (2003) ‘Progressive muscle relaxation and aggression among elementary students with emotional or behavioral disorders.’ Behavioral Disorders, 28, pp. 162–172. https://doi.org/10.1177/ 019874290302800203. Maag, J. (2006) ‘Social skills training for students with emotional and behavioral disorders: A review of reviews.’ Behavioral Disorders, 32 (1), pp. 4–17. https://doi.org/10.1177/019874290603200104. Maggin, D., Briesch, A. & Chafouleas, S. (2013) ‘An application of the What Works Clearinghouse Standards for evaluating single-subject research: Synthesis of the self-management literature base.’ Remedial and Special Education, 34 (1), pp. 44–58. https://doi.org/10.1177/0741932511435176. *Malmgren, K. W., Causton-Theoharis, J. N. & Trezek, B. J. (2005) ‘Increasing peer interactions for students with behavioral disorders via paraprofessional training.’ Behavioral Disorders, 31, pp. 95–106. https://doi.org/10.1177/019874290503100105. Malmgren, K. W. & Meisel, S. M. (2002) ‘Characteristics and service trajectories of youth with serious emotional disturbance in multiple service systems.’ Journal of Child and Family Studies, 11, pp. 217– 229. *Mason, L. H. & Shriner, J. G. (2008) ‘Self-regulated strategy development instruction for writing an opinion essay: Effects for six students with emotional/ behavior disorders.’ Reading & Writing, 21, pp. 71– 93. https://doi.org/10.1007/s11145-007-9065-y. *McDaniel, S., Bruhn, A. & Troughton, L. (2017) ‘A brief social skills intervention to reduce challenging classroom behavior.’ Journal of Behavioral Education, 26, pp. 53–74. https://doi.org/10.1007/ s10864-016-9259-y. *McDaniel, S. C., Houchins, D. E. & Robinson, C. (2016) ‘The effects of check, connect, and expect on behavioral and academic growth.’ Journal of ª 2021 NASEN Emotional and Behavioral Disorders, 24, pp. 42–53. https://doi.org/10.1177/1063426615573262. McGinnis, E. & Goldstein, A. (1997) Skillstreaming the elementary school child: New strategies and perspectives for teaching prosocial skills. Champaign, Illinois: Research Press. McKenna, J. & Brigham, F. (2021) ‘More than de minimis: FAPE in the Post Endrew F. Era.’ Behavior Modification, 45 (1), pp. 3–12. https://doi.org/10.1177/ 0145445519880836. McKenna, J., Flower, A. & Adamson, R. (2015) ‘A systematic review of function-based replacement behavior interventions for students with and at risk for emotional and behavioral disorders.’ Behavior Modification, 40, pp. 678–712. https://doi.org/10.1177/ 0145445515621489. McKenna, J., Flower, A., Falcomata, T. & Adamson, R. (2017) ‘Function-based replacement behavior interventions for students with challenging behavior.’ Behavioral Interventions, 32 (4), pp. 379–398. https:// doi.org/10.1002/bin.1484. McKenna, J., Solis, M., Brigham, F. & Adamson, R. (2019) ‘The responsible inclusion of students receiving special education services for emotional disturbance: Unraveling the practice to research gap.’ Behavior Modification, 43 (4), pp. 587–611. https:// doi.org/10.1177/0145445518762398. *Meyer, K. A. (1999) ‘Functional analysis and treatment of problem behavior exhibited by elementary school children.’ Journal of Applied Behavior Analysis, 32, pp. 229–232. *Miller, N. V., Haas, S. M., Waschbusch, D. A., Willoughby, M. T., Helseth, S. A., Crum, K. I., Coles, E. K. & Pelham, W. E. (2014) ‘Behavior therapy and callous-unemotional traits: Effects of a pilot study examining modified behavioral contingencies on child behavior.’ Behavior Therapy, 45, pp. 606–618. https://doi.org/10.1016/j.beth.2013. 10.006. Mitchell, B., Adamson, R. & McKenna, J. (2017) ‘Curbing our enthusiasm: An analysis of the CheckIn/Check-Out literature using the Council for Exceptional Children’s Evidence-Based Practice Standards.’ Behavior Modification, 41 (3), pp. 343– 367. https://doi.org/10.1177/0145445516675273. Mitchell, B., Kern, L. & Conroy, M. (2019) ‘Supporting students with emotional or behavioral disorders: State of the field.’ Behavioral Disorders, 44 (2), pp. 70–84. https://doi.org/10.1177/0198742918816518. *Peltier, C. & Vannest, K. J. (2018) ‘The effects of schema-based instruction on the mathematical problem solving of students with emotional and behavioral disorders.’ Behavioral Disorders, 43, pp. 277–289. https://doi.org/10.1177/0198742917704647. Poulou, M. (2015) ‘Emotional and behavioral difficulties in preschool.’ Journal of Child and Family Studies, 24 (2), pp. 225–236. 25 Journal of Research in Special Educational Needs,  – Radley, K. C., Dart, E. H., Fischer, A. J. & Collins, T. A. (2020) ‘Publication trends for single-case methodology in school psychology: A systematic review.’ Psychology in the Schools, 57, pp. 683–698. https://doi.org/10.1002/pits.22349. *Rafferty, L. A. (2012) ‘Self-monitoring during whole group reading instruction: Effects among students with emotional and behavioral disabilities during summer school intervention sessions.’ Emotional and Behavioural Difficulties, 17, pp. 157–173. https://doi. org/10.1080/13632752.2012.672866. Rice, E. H. & Yen, C. (2010) ‘Examining gender and the academic achievement of students with emotional disturbance.’ Education and Treatment of Children, 33, pp. 601–621. https://doi.org/10.1353/etc.2010. 0011. Scott, T. & Alter, P. (2017) ‘Examining the case of functional behavior assessment as an evidence-based practice for students with emotional and behavioral disorders in general education classrooms.’ Preventing School Failure, 61 (1), pp. 80–93. https://doi.org/10. 1080/1045988x.2016.1196645. *Scruggs, T. E. & Osguthorpe, R. T. (1986) ‘Tutoring interventions within special education settings: A comparison of cross-age and peer tutoring.’ Psychology in the Schools, 23, pp. 187–193. https:// doi.org/10.1002/1520-6807(198604)23:2<187:aidpits2310230212>3.0.co;2-7. Solis, M., Vaughn, S., Swanson, E. & Mcculley, L. (2012) ‘Collaborative models of instruction: The empirical foundations of inclusion and co-teaching.’ Psychology in the Schools, 49 (5), pp. 498–510. https://doi.org/10.1002/pits.21606. The Peacock Hill Working Group (1991) ‘Problems and promises in special education and related services for children and youth with emotional or behavioral disorders.’ Behavioral Disorders, 16 (4), pp. 299–313. https://doi.org/10.1177/019874299101600406. Travers, J., Cook, B., Therrien, W. & Coyne, M. (2016) ‘Replication research and special education.’ Remedial and Special Education, 37 (4), pp. 195–204. https:// doi.org/10.1177/0741932516648462. *Trussel, R. P., Lewis, T. J. & Stichter, J. P. (2008) ‘The impact of targeted classroom and function-based behaveor interventions on problem behaviors of students with emotional/behavioral disorders.’ Behavioral Disorders, 33, pp. 153–166. https://doi. org/10.1177/019874290803300303. U.S Department of Education, EDFacts Data Warehouse (2017) IDEA part B child count and educational environments collection 2016–2017. <https:// www2.ed.gov/programs/osepidea/618-data/static-table s/index.html>. Vinh, M., Strain, P., Davidon, S. & Smith, B. J. (2016) ‘One state’s systems change efforts to reduce child 26 care expulsion: Taking the Pyramid Model to scale.’ Topics in Early Childhood Special Education, 36 (3), pp. 159–164. https://doi.org/10.1177/ 0271121415626130. Wakschlag, L., Briggs-Gowan, M., Carter, A., Hill, C., Danis, B., Keenan, K., McCarthy, K. & Leventhal, B. (2007) ‘A developmental framework for distinguishing disruptive behavior from normative misbehavior in preschool children.’ Journal of Child Psychology and Psychiatry, 48, pp. 976–987. https://doi.org/10.1111/j. 1469-7610.2007.01786.x. Walker, H. M., Ramsey, E. & Gresham, F. M. (2004) Antisocial Behavior in School: Evidence-based Practices (2nd edn). Belmont, CA: Thomson/ Wadsworth. Watts, G., Bryant, D. & Carroll, M. (2019) ‘Students with emotional-behavioral disorders as cross-age tutors: A synthesis of the literature.’ Behavioral Disorders, 44 (3), pp. 131–147. https://doi.org/10. 1177/0198742918771914. *Weeden, M., Wills, H. P., Kottwitz, E. & Kamps, D. (2016) ‘The effects of a class-wide behavior intervention for students with emotional and behavioral disorders.’ Behavioral Disorders, 42, pp. 285–293. https://doi.org/10.17988/BD-14-12.1. *Wehby, J. H., Falk, K. B., Barton-Arwood, S., Lane, K. L. & Cooley, C. (2003) ‘The impact of comprehensive reading instruction on the academic and social behavior of students with emotional and behavioral disorders.’ Journal of Emotional and Behavioral Disorders, 11, pp. 225–238. https://doi. org/10.1177/10634266030110040401. *Wehby, J. H., Lane, K. L. & Falk, K. B. (2005) ‘An inclusive approach to improving early literacy skills of students with emotional and behavioral disorders.’ Behavioral Disorders, 30, pp. 155–169. https://doi. org/10.1177/019874290503000208. Williams, J., Miciak, J., McFarland, L. & Wexler, J. (2016) ‘Learning disability identification criteria and reporting in empirical research: A review of 2001– 2013.’ Learning Disabilities Research & Practice, 31, pp. 221–229. https://doi.org/10.1111/ldrp.12119. Withey, K. (2018) ‘Interventions for young children with and at risk for emotional and behavioral disorders.’ Intervention in School and Clinic, 53 (3), pp. 183– 187. https://doi.org/10.1177/1053451217702110. Yell, M. (2019) ‘Endrew F. v. Douglas County School District (2017): Implications for educating students with emotional and behavioral disorders.’ Behavioral Disorders, 45 (1), pp. 53–62. https://doi.org/10.1177/ 0198742919865454. Zabel, R., Kaff, M. & Teagarden, J. (2013) ‘Promoting interdisciplinary practice: An interview with Steven R. Forness.’ Intervention in School and Clinic, 49, pp. 255–260. ª 2021 NASEN