Next Article in Journal
A Comparative Analysis of Lung Cancer Incidence and Tobacco Consumption in Canada, Norway and Sweden: A Population-Based Study
Previous Article in Journal
Predictors of Unmet Needs in Chilean Older People with Dependency: A Cross-Sectional Study
Previous Article in Special Issue
Impact of COVID-19 Pandemic on Children’s Fundamental Motor Skills: A Study for the Taiwanese Preschoolers Teachers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Active School Commuting in School Children: A Narrative Review of Current Evidence and Future Research Implications

School of Health Sciences, College of Health and Medicine, University of Tasmania, Launceston, TAS 7250, Australia
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2023, 20(20), 6929; https://doi.org/10.3390/ijerph20206929
Submission received: 11 September 2023 / Revised: 6 October 2023 / Accepted: 11 October 2023 / Published: 16 October 2023
(This article belongs to the Special Issue Child Physical Activity and Health)

Abstract

:
Active school commuting (ASC) has been proposed as a practical way to inculcate positive physical activity habits in children. This paper reviews the current evidence regarding ASC among children, highlights advances in research techniques and existing limitations in the field, and outlines future implications for research and promotion. A comprehensive literature search was conducted to identify English language studies on ASC among children aged 6–12 years, followed by a narrative review. ASC has witnessed a global decline, despite evidence of its contribution to physical activity levels. Context-dependent factors such as commuting distance and parental safety concerns are consistently identified as key determinants of ASC. Several promising interventions have been identified. Despite the limitations in intervention scope and quality, notable advancements in research techniques, such as multilevel regression and agent-based modelling, have been identified. Effective promotion of ASC to tackle childhood physical inactivity requires collaborative efforts among schools, parents, and the government, and should be tailored to address multilevel determinants within the local context. Future research should leverage recent advancements in research techniques to develop effective promotion strategies, while considering the context-dependent nature of ASC behaviours and addressing existing limitations, including the lack of standardised definitions and limited geographical and age coverage.

1. Introduction

Childhood physical inactivity continues to pose a significant public health challenge, with a majority of children and adolescents failing to meet the World Health Organization’s physical activity recommendations [1]. Unfortunately, the COVID-19 pandemic has likely exacerbated this issue due to social distancing restrictions and reduced access to physical activity facilities such as school, sport clubs, and swimming pools [2]. Given the propensity for habits formed in early life to transition into adulthood [3,4,5], early and middle childhood provides an opportune window to inculcate positive physical activity habits that can potentially manifest into life-long health benefits [6].
Active school commuting (ASC) was first proposed in the early 2000s as a potential source of physical activity for school children beyond the school boundaries [7,8]. With each school child typically making 360–400 trips to and from school each year [9], these journeys present an excellent opportunity to integrate physical activity into their daily routines, especially through active modes of transportation such as walking, cycling, or using non-motorised vehicles [10]. The past two decades have witnessed ASC emerging as a key area of childhood physical activity research. Nevertheless, the fact that childhood physical inactivity remains entrenched [11] makes it imperative at this juncture to examine the roles of ASC in promoting physical activity and to identify strategies to maximise its contribution in fostering a healthier and more active childhood.
This review aims to provide an overview of the current evidence in ASC. It also highlights advances in research techniques and existing limitations in the field, followed by a discussion of implications for future research and promotion.

2. Materials and Methods

This narrative review focuses on school children aged between 6 and 12 years, the typical age range of primary or elementary school students globally [12]. It seeks to provide a brief overview of the following key issues pertaining to ASC:
  • What are the trends of the global prevalence of ASC?
  • How does ASC contribute to the physical activity levels of school children?
  • What are the health benefits of ASC?
  • What are the determinants of ASC behaviours?
  • What interventions are effective in increasing ASC?
  • What are the advancements in research techniques related to ASC?
  • What are the current limitations in ASC research?
A literature search was conducted in March 2023 across five electronic databases (PubMed, Scopus, ScienceDirect, Transport Research International Documentation (TRID), and Google Scholar) using a combination of search terms (Table 1). Additional literature was identified by reviewing the reference lists of identified articles. Furthermore, searches for ongoing studies and study protocols were conducted at trial registries, including the International Clinical Trials Registry Platform (ICTRP) and ClinicalTrials.gov. After removal of duplicates, articles were screened using the inclusion and exclusion criteria (Table 2). The full texts of the remaining papers were then reviewed and assigned under the relevant key issues. The search process is illustrated in Figure 1.
During the review, emphasis was placed on findings from systematic reviews or synthesis, supplemented by individual studies subsequently published. If a systematic review addressing a specific issue was not available, or if the existing systematic reviews did not provide sufficient information, findings from individual studies were reviewed and discussed. A total of 158 articles were included in the final review (Table S1).

3. Results

3.1. What Are the Trends of the Global Prevalence of ASC?

Before the turn of the millennium, national longitudinal data on how school children commute to school was only available from travel behaviour surveys in a few countries, including New Zealand [13], Switzerland [14], the United Kingdom [15], and the United States [16]. However, with the growing recognition of the importance of ASC, countries like Australia [17], the Czech Republic [18], Finland [19,20,21], and Spain [22] started collecting such data as part of their national health surveys or census since the mid-2000s. Often these datasets are also supplemented by state- or city-based surveys conducted [23,24,25,26].
Taken together, the available data reveal a significant and consistent decline in the prevalence of ASC from the 1970s to the early 2000s, with only a handful of countries showing signs of plateauing reductions or improvement (Figure 2). As the decline in ASC has been observed across developed and developing countries, some researchers have considered it a global phenomenon [27].

3.2. How Does ASC Contribute to the Physical Activity Levels of School Children?

Numerous studies have explored the impact of ASC on physical activity levels in school children with various methods including accelerometers, pedometers, physical activity diaries, or self-reported questionnaires. Systematic reviews of these studies concluded that ASC is associated with increased total daily physical activity levels [30,31,32,33]. The majority of evidence comes from cross-sectional studies which gathered the information at one point in time. Accelerometry-based studies reported that actively commuting children accumulate an additional 3.3 to 40 min of moderate-to-vigorous physical activity (MVPA) compared to their passively commuting counterparts [34,35,36,37,38,39,40,41,42,43,44,45,46]. Similarly, pedometer-based studies have revealed that walking to and from school is associated with higher daily step counts that range from 1000 to 3400 steps [47,48,49,50,51,52,53,54,55]. Furthermore, diary-based or questionnaire studies indicated that actively commuting children are more likely to meet requirements from physical activity guidelines [56,57,58,59].
On the other hand, among the few longitudinal studies which compared the physical activity level of actively and passively commuting children across different time points, the results were mixed, with a few reporting significant differences favouring ASC at either baseline or follow-up within specific sex or age groups [60,61,62,63]. Two Swedish studies reported no difference at both baseline and follow-up at 1 and 2 years [64,65]. While reasons behind these inconsistencies were not explored, one longitudinal study reported a noteworthy finding that a switch from passive to active commuting was associated with an increase in daily minutes of MVPA [66]. This was further supported by several interventional studies on walking school bus programmes, most of which have reported increases in physical activity level for school children assigned to walk to or from their schools [67,68,69].
Most of the above studies combined walking and cycling as one single “active commuting” category in their data collection and analysis. Only a small number of cross-sectional studies compared the level of physical activity between children who walked and cycled and reported inconsistent findings. One accelerometer-based study in Denmark found that children who walked to school recorded higher values for MVPA than those who cycled, but the differences were not statistically significant [70]. In contrast, one questionnaire-based study in the United Kingdom found that cycling is associated with higher physical activity scores than walking [71].
It has been proposed that the contribution of ASC to the daily total physical activity level can operate through two mechanisms: direct and indirect [31]. The direct mechanism involves children gaining physical activity directly through the active commuting itself, while the indirect mechanism encompasses a spill-over effect, where ASC fosters a positive attitude toward physical activity, leading children to seek out other opportunities for active play and exercise outside the commuting journeys. The presence of the direct mechanism is supported by several observations. Firstly, the differences in physical activity levels between actively and passively commuting children were found to be more pronounced on weekdays and less noticeable on weekends, suggesting that ASC contributes significantly to daily physical activity on school days [35,72]. Secondly, most studies comparing physical activity levels during the school commuting journeys consistently found differences between actively and passively commuting children. In contrast, the evidence pertaining to the indirect mechanism is less conclusive, with most studies being unable to identify significant differences in physical activity levels during school hours or free time between children who actively commute and those who do not [35,38,57,73,74,75,76,77]. These findings suggest that while ASC contributes positively to physical activity level, this impact is confined only to the school commuting journeys and may not translate to contexts outside these time periods.

3.3. What Are the Health Benefits of ASC?

Numerous studies have examined the health benefits of ASC in children, with mixed findings across various health indicators. Most were of cross-sectional design which measured children at a single point in time. Regarding adiposity measures. while some cross-sectional studies demonstrated associations with lower body mass index (BMI) and skin fold thickness in actively commuting children, most longitudinal studies, with follow-up periods ranging from 10 weeks to 6 years, did not observe the same beneficial effect [30,31,32,33,37,60,63,78,79,80,81,82,83,84]. Cross-sectional studies which directly compared the BMI between children who walk and cycle to school were also unable to detect any significant between-group difference [46,71,85,86,87]. Similarly, studies measuring waist circumference [37,52,62,63,81,88,89,90,91] and those that utilised more reliable assessments of body composition, such as dual-energy X-ray absorptiometry [64,65], bioimpedance analysis [81,92], and air displacement plethysmography [75], did not find any discernible difference between active and passive school commuters.
Meanwhile, studies have consistently shown that children who cycle to school have a higher cardiovascular fitness level when compared with non-cyclists, including those who commute by walking or by car [70,87,93,94,95,96]. However, this relationship is not as evident between active and passive school commuters [63,71,81,88,89,97,98,99,100,101]. Findings for other measures of muscular fitness, such as hand-grip strength and standing jump, have also been inconsistent [63,87,92,101,102,103,104].
Most studies on cognitive skills reported no significant differences between active and passive school commuters [81,105,106], except for better spatial cognition in children who walked to school [107,108,109]. Similarly, studies on academic achievement are limited and have yielded mixed results [110,111,112,113]. On the other hand, ASC has been associated with fewer depressive symptoms [90] and better psychological well-being [114]. Studies on other health benefits, such as bone health and risk indicators for metabolic syndrome and cardiovascular diseases, have not yielded conclusive results [62,64,65,115,116,117].

3.4. What Are the Determinants of ASC Behaviours?

Like many other health behaviours, ASC is influenced by a variety of factors. While the objective built environment was traditionally considered the primary determinant [7], research adopting behavioural approaches, such as the social ecological model, has identified additional critical determinants from individual, social, and policy levels [118,119]. McMillan’s framework of the Elementary-Aged Child’s Travel Behaviour was the first to address active school commuting in school children. While acknowledging the link between objective built environment and school commuting, this framework emphasised the role of parents as the principal decisionmakers regarding the active transport behaviour of children. As such, parents’ perceptions of the commuting environment and household transportation options during the decision-making process cannot be underestimated [120] (Figure 3).
The impact of these determinants varies significantly across different study locations. Nevertheless, some consistent findings have emerged from the literature [119,121,122,123,124,125,126,127] as outlined in Table 3. For instance, positive associations were found between active commuting and environmental features like walkability and presence of physical activity facilities, while longer school journey distance and higher household car ownership had negative effects. Parental concerns regarding crime and traffic safety also play a critical role in shaping commuting behaviour. Several sociodemographic factors, including ethnicity, parental education, income levels, and neighbourhood socioeconomic status, were also associated with ASC behaviour.
Two systematic reviews on parental barriers of ASC have identified several factors that parents consider during the decision-making process, such as the lack of time to accompany the children to walk or cycle to school, ease of dropping off children on the way to work by car, and heaviness of their children’s backpacks [119,124]. Nevertheless, most of these findings were from the United States, thus limiting their generalisation. Qualitative studies have provided further insights into the parental decision-making process. Parents weighed the benefits of active commuting against safety concerns, distance, and practicality [128,129,130,131,132], while children’s perspectives highlighted the negotiation of risks, enjoyment, social interactions, and exploration during the journey [133,134].
On the other hand, the role of school in facilitating ASC appears to be under-explored in the existing literature. A systematic review of parental perceived barriers has identified four studies conducted in the United States which showed that parental perception of encouragement of ASC by school and availability of a school bus service were positively and negatively associated with ASC, respectively [124,135,136,137,138].
Table 3. Key determinants of ASC behaviours.
Table 3. Key determinants of ASC behaviours.
Type of EvidenceCategoriesFactors IdentifiedKey FindingsReferences
QuantitativeObjective built environmentWalkabilityHigher neighbourhood walkability is positively associated with ASC[121,122,126]
Residential densityHigher residential density is positively associated with walking to school[121,122]
Land use mixIncreased land use mix diversity is positively associated with walking to school[121,122]
Density of physical activity facilities in the neighbourhoodAssociated with a higher rate of ASC among school children[121,122]
LocalityNeighbourhoods located in cities or urban regions have a higher rate of ASC compared with their rural counterparts[121]
Household transportation optionsDistance of school journeyIncreased distance negatively correlated with ASC; the distance threshold varies depending on population and country[123,125,126]
Car ownership within the householdAssociated with a decreased likelihood of children engaging in ASC[121]
Perceived commuting environment by parentsPerceived distance of the school journeyIdentified as a major obstacle to active commuting[119,123,124]
Traffic safety concerns (e.g., heavy traffic, fast-moving vehicles, unsafe driving behaviours)Associated with lower odds of active commuting to school[119,123,124]
Crime-related safety concernsAssociated with lower odds of active commuting to school[119,123,124]
Social support (presence of other children and adults) along the commute routeAssociated with higher odds of active commuting to school[124]
Perceived neighbourhood walkabilityAssociated with higher odds of active commuting to school[124]
Discrepancy between subjective perception and objective realityLack of correlation between parents’ perception of neighbourhood safety and actual crime and traffic crash rate[136,139]
Sociodemographic factorsSexASC is less prevalent in girls than in boys, with boys more likely to cycle to school while girls prefer walking[140]
AgeConflicting results on the association between ASC and the age of the child, with most studies demonstrating null effect[141]
Household incomeChildren from families with higher household income are less likely to engage in active commuting to school[125]
EthnicityChildren from non-white ethnic groups are less likely to engage in active commuting to school[125]
Parental educationChildren of parents with higher levels of education are less likely to actively commute to school[125]
Family structureChildren with a single parent are less likely to commute actively to school, while those who are only children are more likely to do so[121]
Neighbourhood socioeconomic statusNeighbourhood with a higher socioeconomic status is weakly positively associated with a higher rate of ASC[127]
Qualitative Parents’ decision making on mode of school commuteParents’ attitudes towards ASC are shaped by a negotiation process. They balance the benefits of active commuting with concerns about traffic and their children’s safety, along with other factors like distance to school, child’s maturity, and time constraints
Regular ASC by children reduces parents’ perceived barriers
[128,129,130,131,142]
Children’s perspectiveChildren may feel vulnerable during the ASC journey but are able to manage the risks. They find the commute journey enjoyable as it allows for social interactions and exploration of the environment[134]

3.5. What Interventions Are Effective in Increasing ASC?

A growing body of studies have investigated the effectiveness of various types of intervention, including safe routes to school (SRTS) [7,143], walking school bus (WSB) [144], bicycle trains, and school travel plans [145], on promoting ASC among school children. As assessed by several systematic reviews, most of these interventions can result in increased ASC and physical activity among children, albeit with small effect sizes [144,146,147,148,149,150,151]. Nevertheless, their benefits can be significant at the population level [148], and economic analyses have further supported their cost-effectiveness [152].
Although the evidence is limited, several key factors may have contributed to the success of these interventions. Interventions which incorporate both educational activities and infrastructure modifications [148], and those with a stronger involvement from schools and parents, have shown greater effectiveness [146]. Interventions which targeted ASC instead of physical activity in general are also more effective [146]. These findings are consistent with the principles of the social ecological model, which emphasises the importance of addressing key factors from multiple levels to create an enabling environment.
Several barriers that may threaten their effectiveness have also been identified. Many studies reported substantial variation of implementation across schools due to the varied degrees of motivation among schools [148]. The limited availability of funding and resources also affects the long-term sustainability of the interventions [148]. Furthermore, some initiatives which depend on volunteer participation, such as WSB programmes, face on-going difficulties in recruiting sufficient volunteers [144].

3.6. What Are the Advancements in Research Techniques Related to ASC?

Increasing utilisation of geospatial data: One notable development is the increased utilisation of geographic information systems (GISs) in measuring the objective built environment. Prior to the advent of GISs, time- and resource-intensive on-site auditing was the primary method for obtaining such information [126]. In recent years, spatial datasets of street networks, land use information, and locations of schools and physical activity facilities have become more accessible from government and non-profit organisations alongside the rapid advancement of a variety of GIS software which has increased the processing and analysis of these data drastically [153]. Furthermore, these spatial datasets can be combined with area- or neighbourhood-level data from other sources, such as census data, to generate secondary spatial datasets that capture crucial factors found to be positively associated with ASC [126]. One notable example concerns the measurement of school commute distance. While early studies relied on parental reports that are prone to inaccuracies [154], the availability of GIS enables a more precise estimation by calculating straight-line distances between home and school using spatial data [155]. However, this approach does not account for the actual commuting route taken. To address this limitation, subsequent studies have incorporated routing algorithms, such as those from ArcGIS and Google Maps, to estimate the shortest possible routes [85,156]. More recently, more accurate methods have been adopted, which include having school children draw their commuting routes on an electronic map [157] or wear GPS monitors to capture their actual route [158,159].
Increasing utilisation of standardised instruments: Another advancement involves the increasing use of standardised instruments which has enabled direct comparisons across studies. One notable example is the Neighbourhood Environment Walkability Scale (NEWS), a questionnaire designed to measure residents’ perceptions of the environmental attributes of their local area in relation to physical activity [160]. NEWS and its derivatives [161,162] have been employed in studies on ASC to gauge parents’ and children’s perception on their neighbourhood’s walkability [85,163,164,165,166]. Recently, there have also been initiatives to develop and validate standardised instruments to assess and measure school children’s ASC behaviours. Examples include the Mode and Frequency of Commuting To and From School Questionnaire [167,168] and its derivative, the Family Commuting-to-School Behaviour Questionnaire [169].
Advancements in analytical approaches: Early studies predominantly employed simple univariate or bivariate statistical techniques at the individual level, overlooking the multilevel nature of the interactions [119] as well as the social and spatial nesting of school children [170]. Inspired by McMillan’s framework, a number of theoretical frameworks on active school commuting, grounded in the social ecological model, have emerged in recent years [118]. These frameworks offer diverse perspectives on how parents’ decision-making processes are influenced by different factors across various levels. The availability of these different frameworks has supported the adoption of more sophisticated analytical methods, such as multilevel regression modelling, structural equation modelling, and hierarchical modelling as seen in some recent studies [157,170,171,172,173]. These approaches account for the nested and multilevel relationships between school children, parents, neighbourhoods, and schools, thus enabling more appropriate data analysis and a deeper understanding of the complex and interrelated factors that influence school commuting behaviour.
Emerging application of modelling techniques: The logistical complexity and resource requirements of evaluating ASC interventions under real-world settings have prompted researchers to adopt alternative modelling approaches. The logic pathway modelling to evaluate the national effects of a walking school bus intervention on body mass index and disability-adjusted life years is a good case in point [174]. Agent-based modelling (ABM) is another modelling technique gaining popularity in recent years [175]. ABM explicitly models the decision-making processes of individual parents and children regarding school commute modes and incorporates key determinants such as distance, safety concerns, and social norms. This approach allows researchers to simulate in silico different intervention scenarios and predict their impact on ASC behaviours, thus providing valuable insights for intervention design and evaluation. Several studies have utilised ABM to study ASC and demonstrated its potential for guiding future research in this field [176,177,178,179].

3.7. What Are the Current Limitations in ASC Research?

Classification of ASC: The lack of standardised classification schemes for ASC makes it difficult to compare and generalise findings across studies and localities [180]. The widely varying thresholds for classifying school children as active commuters, as well as the combination of walking and cycling as a single “active commuting” category, may have weakened the findings with determinants, physical activity levels, and health benefits and effective interventions [30]. Notably, studies that measured ASC as an ordinal or continuous variable, and those which separated walking and cycling, tend to report significant findings [36,47,59].
Limited age and geographical coverage: Most ASC studies focus on older children aged nine years or above, with limited representation of younger age groups. Extrapolation of findings from older school children might not be appropriate, as younger children may have different commuting patterns and constraints. For instance, they have lower level of independent mobility [33], may live closer to their schools, and will only walk when the school is nearby or with adult supervision [32]. Moreover, research on ASC is predominantly derived from metropolitan or urban areas in developed nations, with limited data from rural, regional, and developing countries [78,147].
Heterogeneity, mediators, and confounders in study methodology: The substantial heterogeneity among studies in terms of design (e.g., cross-sectional, longitudinal, non-randomised, and randomised controlled trials), definitions, and methodology has impeded the conduct of meta-analysis. Consequently, most systematic reviews have to rely on the questionable vote-counting technique for summarising results [181]. The wide variation of study findings across different cultures, economies, and geographical locations attests to the importance of contextual factors in shaping school commuting behaviours [122]. However, the effects of various mediators and confounders, including diet, school journey distance, ethnicity, culture, and socioeconomic status are often overlooked in existing studies [148].
Limited research in key areas: The scarcity of studies on the benefits of ASC, such as mental health and academic performance, may contribute to the limited appeal of ASC among parents and policymakers [30,31,32,33,78,79,80,104,182,183]. Similarly, the existing body of qualitative literature on this topic is also limited [134], thus stymieing a more comprehensive understanding of the dynamics underlying ASC behaviours. Interventional studies also suffer from weak quality resulting from non-randomised design, lack of blinding and control groups, and reliance on self-reporting [147,148,149]. Moreover, there is also a lack of utilisation of behavioural frameworks in interventional studies, which has limited the understanding on the mechanism underlying behavioural changes and the generalisability of findings to other settings [146,147,148].

4. Discussion

4.1. Future Research Implications

Future research should aim to adopt a clearer and standardised definition of ASC, while also analysing different forms of active commuting separately [119] and as ordinal or continuous variables [180]. Researchers should broaden their study scope by targeting younger age groups and exploring ASC behaviours in rural, regional, and developing areas. They should go beyond cross-sectional studies by utilising longitudinal studies with extended follow-up periods to better understand the long-term effects on physical activity level and other health indicators. Moreover, like several recent studies [184,185], future studies may wish to adopt a mixed-methods approach, which combines quantitative and qualitative techniques [186] to generate more useful information regarding the local context within a single study. Given the complex and context-specific nature of ASC behaviours, the use of more sophisticated analytical methods, such as multilevel regression modelling and structural equation modelling, can help elucidate the effects of various mediators and confounders. This in turn can inform the development of targeted interventions and policies that address the specific needs and challenges of the local context.
Future interventional studies should focus on optimisation of study quality to generate robust evidence. This can be achieved through the utilisation of representative samples, larger sample sizes, randomised controlled trial designs, and the use of valid and reliable objective measures [148,150,151]. In addition, interventional studies should be grounded in a theoretical framework and incorporate mediators identified from behavioural studies. Furthermore, longer follow-up periods are necessary to evaluate the long-term efficacy of interventions. To overcome resource limitations, modelling techniques, such as agent-based modelling, can be utilised to plan and identify potentially effective interventions.

4.2. Implications for Promotion of ASC

Promoting ASC requires a comprehensive and evidence-based approach at both the national and local levels. Despite some mixed findings, the existing evidence lends support to the positive impact of ASC on increasing physical activity levels among school children. Given the current urgency of the childhood physical inactivity and obesity crisis, ASC should be recognised as a pragmatic response and an integral part of the school health policy.
To ensure their effective implementation, it is important to address the lack of prevalence data on ASC in many regions of the world. Establishing regular surveillance of ASC prevalence through national or regional health surveys becomes imperative. Such data will not only inform stakeholders about the effectiveness of current initiatives but also shape future strategies. In the absence of such a mechanism, a baseline situational analysis of the ASC in the target region is indispensable.
Schools play a pivotal role in fostering ASC habits among students. To promote ASC effectively, a context-specific approach is required at the school level. Local partnerships between government, schools, parents, and community members allow tailoring of strategies to the specific needs and challenges of their communities. While reducing school commuting distances for existing schools may be impractical [123], a multifaceted approach that includes infrastructure enhancements, provision of safe and supportive commuting environments, and educational initiatives can positively influence other modifiable determinants, such as parental safety concerns. Recognising the role of parents as major decisionmakers, highlighting the known benefits of ASC, such as improved fitness and mental health, can help counteract some of the parental barriers. Collaborative efforts are necessary to also overcome general and locally specific challenges, including resource limitations.
Equity considerations are also important to ensure all school children have access to ASC opportunities. Given the positive correlation between neighbourhood socioeconomic status, crime safety concerns, and ASC prevalence, targeted interventions that address transportation inequities and creating safe routes for walking and cycling in disadvantaged areas are needed. Nevertheless, the inverse relationship between ASC and income levels implies the possible existence of a gradient of ASC preference even in wealthier neighbourhoods. Hence, strategies are also needed to focus on the specific factors influencing ASC in different families.

5. Conclusions

This review offers a comprehensive overview of the current landscape of ASC research, highlighting its potentials in addressing childhood physical inactivity (Figure 4). Current evidence supports the direct contribution of ASC to physical activity levels and several health benefits. The decline in global prevalence of ASC presents an opportunity for it to be one of the strategies to promote regular physical activity among school children. Behavioural frameworks encompassing multilevel determinants have facilitated the understanding of ASC behaviours. Several promising interventions, such as the walking school bus, have also be identified. While advances in research techniques have supported the progress in ASC research, it is essential to address the existing limitations and challenges in methodology and implementation, including the lack of standardised ASC definition and paucity of research in younger age groups and rural settings.
Although future research aimed at addressing the identified research gaps is expected to be carried out mostly in developed countries, it is crucial to acknowledge that ASC remains under-examined in many developing countries and rural areas of developed countries. As ASC behaviour is highly context specific, it is prudent for researchers to first study the local situation and comprehend the local dynamics contributing to the observed ASC behaviours, using quantitative and qualitative study methods supported by relevant behavioural frameworks. This approach can help identify critical local barriers and enablers of ASC behaviour, on which targeted interventions can be developed in partnerships involving the government, schools, parents, and local community. Modelling using ABM can be considered to select the interventions with the highest potential for effectiveness.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijerph20206929/s1, Table S1: Information of papers included in the review.

Author Contributions

Conceptualisation: H.Y.L., S.J., K.D.K.A. and A.P.H.; Methodology: H.Y.L.; Formal analysis: H.Y.L.; Investigation: H.Y.L.; Writing—Original Draft: H.Y.L.; Writing—Review and Editing: S.J., K.D.K.A. and A.P.H.; Visualisation: H.Y.L.; Supervision: S.J., K.D.K.A. and A.P.H.; Project administration: H.Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

H.Y.L. was supported by a Tasmania Graduate Research Scholarship and a tuition fee scholarship awarded by the University of Tasmania.

Institutional Review Board Statement

Ethical review and approval were waived for this study since the preparation of this paper did not involve primary research or data collection with human subjects.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. World Health Organization. Global Status Report on Physical Activity 2022; World Health Organization: Geneva, Switzerland, 2022. [Google Scholar]
  2. Neville, R.D.; Lakes, K.D.; Hopkins, W.G.; Tarantino, G.; Draper, C.E.; Beck, R.; Madigan, S. Global Changes in Child and Adolescent Physical Activity during the COVID-19 Pandemic: A Systematic Review and Meta-analysis. JAMA Pediatr. 2022, 176, 886–894. [Google Scholar] [CrossRef] [PubMed]
  3. Telama, R.; Yang, X.; Laakso, L.; Viikari, J. Physical Activity in Childhood and Adolescence as Predictor of Physical Activity in Young Adulthood. Am. J. Prev. Med. 1997, 13, 317–323. [Google Scholar] [CrossRef] [PubMed]
  4. Telama, R.; Yang, X.; Viikari, J.; Välimäki, I.; Wanne, O.; Raitakari, O. Physical activity from childhood to adulthood: A 21-year tracking study. Am. J. Prev. Med. 2005, 28, 267–273. [Google Scholar] [CrossRef] [PubMed]
  5. Boreham, C.; Riddoch, C. The physical activity, fitness and health of children. J. Sports Sci. 2001, 19, 915–929. [Google Scholar] [CrossRef] [PubMed]
  6. Centers for Disease Control and Prevention. Physical Activity Facts. Available online: https://www.cdc.gov/healthyschools/physicalactivity/facts.htm (accessed on 20 March 2023).
  7. Tudorlocke, C.; Ainsworth, B.E.; Popkin, B.M. Active Commuting to School: An overlooked source of childrens’ physical activity? Sports Med. 2001, 31, 309–313. [Google Scholar] [CrossRef] [PubMed]
  8. Hills, A.P.; Cambourne, B. Walking to school—A sustainable environmental strategy to prevent childhood obesity. Australas. Epidemiol. 2002, 9, 2. [Google Scholar]
  9. Craw, J. Statistic of the Month: How Much Time do Students Spend in School? Available online: https://ncee.org/quick-read/statistic-of-the-month-how-much-time-do-students-spend-in-school/ (accessed on 20 March 2023).
  10. Tilse, M.; Herriot, M. 72 Be Active Australia: A framework for health sector action for physical activity 2005–2010. J. Sci. Med. Sport 2005, 8, 47. [Google Scholar] [CrossRef]
  11. World Health Organization. Global Action Plan on Physical Activity 2018–2030: More Active People for a Healthier World; World Health Organization: Geneva, Switzerland, 2018. [Google Scholar]
  12. UNICEF Data. Primary Education. Available online: https://data.unicef.org/topic/education/primary-education/ (accessed on 20 March 2023).
  13. New Zealand Household Travel Survey. HD011—Mode Share of Journeys to School, Aged 5–12. Available online: https://www.transport.govt.nz/statistics-and-insights/household-travel/sheet/why#element-278 (accessed on 23 March 2023).
  14. Grize, L.; Bringolf-Isler, B.; Martin, E.; Braun-Fahrländer, C. Trend in active transportation to school among Swiss school children and its associated factors: Three cross-sectional surveys 1994, 2000 and 2005. Int. J. Behav. Nutr. Phys. Act. 2010, 7, 28. [Google Scholar] [CrossRef]
  15. Black, C.; Collins, A.; Snell, M. Encouraging Walking: The Case of Journey-to-school Trips in Compact Urban Areas. Urban Stud. 2016, 38, 1121–1141. [Google Scholar] [CrossRef]
  16. Mcdonald, N. Active Transportation to School: Trends among U.S. Schoolchildren, 1969–2001. Am. J. Prev. Med. 2007, 32, 509–516. [Google Scholar] [CrossRef]
  17. Australian Bureau of Statistics. Census at School Australia—National Time Series Tables. Available online: http://www.abs.gov.au/websitedbs/CaSHome.nsf/Home/CensusAtSchool%2BData%2BTime%2BSeries (accessed on 23 March 2023).
  18. Pavelka, J.; Sigmundová, D.; Hamřík, Z.; Kalman, M.; Sigmund, E.; Mathisen, F. Trends in Active Commuting to School among Czech Schoolchildren from 2006 to 2014. Central Eur. J. Public Health 2017, 25 (Suppl. S1), S21–S25. [Google Scholar] [CrossRef]
  19. LIKES Research Centre for Physical Activity and Health. Finland’s Report Card 2016 on Physical Activity for Children and Youth; LIKES Research Centre for Physical Activity and Health: Jyväskylä, Finland, 2016. [Google Scholar]
  20. LIKES Research Centre for Physical Activity and Health. Finland’s Report Card 2018 on Physical Activity for Children and Youth; LIKES Research Centre for Physical Activity and Health: Jyväskylä, Finland, 2018. [Google Scholar]
  21. JAMK University of Applied Sciences. Finland’s Report Card 2022 on Physical Activity for Children and Youth; JAMK University of Applied Sciences: Jyväskylä, Finland, 2022. [Google Scholar]
  22. Gálvez-Fernández, P.; Herrador-Colmenero, M.; Esteban-Cornejo, I.; Castro-Piñero, J.; Molina-García, J.; Queralt, A.; Aznar, S.; Abarca-Sos, A.; González-Cutre, D.; Vidal-Conti, J.; et al. Active commuting to school among 36,781 Spanish children and adolescents: A temporal trend study. Scand. J. Med. Sci. Sports 2021, 31, 914–924. [Google Scholar] [CrossRef] [PubMed]
  23. van der Ploeg, H.P.; Merom, D.; Corpuz, G.; Bauman, A.E. Trends in Australian children traveling to school 1971–2003: Burning petrol or carbohydrates? Prev. Med. 2008, 46, 60–62. [Google Scholar] [CrossRef] [PubMed]
  24. de Sá, T.H.; Garcia, L.M.T.; Mielke, G.I.; Rabacow, F.M.; de Rezende, L.F.M. Changes in travel to school patterns among children and adolescents in the São Paulo Metropolitan Area, Brazil, 1997–2007. J. Transp. Health 2015, 2, 143–150. [Google Scholar] [CrossRef]
  25. Buliung, R.N.; Mitra, R.; Faulkner, G. Active school transportation in the Greater Toronto Area, Canada: An exploration of trends in space and time (1986–2006). Prev. Med. 2009, 48, 507–512. [Google Scholar] [CrossRef] [PubMed]
  26. Cui, Z.; Bauman, A.; Dibley, M.J. Temporal trends and correlates of passive commuting to and from school in children from 9 provinces in China. Prev. Med. 2011, 52, 423–427. [Google Scholar] [CrossRef] [PubMed]
  27. Yang, Y.; Xue, H.; Liu, S.; Wang, Y. Is the decline of active travel to school unavoidable by-products of economic growth and urbanization in developing countries? Sustain. Cities Soc. 2019, 47, 101446. [Google Scholar] [CrossRef] [PubMed]
  28. Woods, C.; Powell, C.; Saunders, J.A.; O’Brien, W.; Murphy, M.H.; Duff, C.; Farmer, O.; Johnston, A.; Connolly, S.; Belton, S.; et al. The Children’s Sport Participation and Physical Activity Study 2018 (CSPPA 2018); Sport Ireland: Dublin, Ireland, 2019. [Google Scholar]
  29. UK Department for Transport. NTS0615: Usual Mode of Travel to School by Age Group, England: 1995 to 1997 Onwards. Available online: https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file/1101119/nts0615.ods (accessed on 24 March 2023).
  30. Larouche, R.; Saunders, T.J.; Faulkner, G.E.J.; Colley, R.; Tremblay, M. Associations Between Active School Transport and Physical Activity, Body Composition, and Cardiovascular Fitness: A Systematic Review of 68 Studies. J. Phys. Act. Health 2014, 11, 206–227. [Google Scholar] [CrossRef]
  31. Lee, M.C.; Orenstein, M.R.; Richardson, M.J. Systematic Review of Active Commuting to School and Children’s Physical Activity and Weight. J. Phys. Act. Health 2008, 5, 930–949. [Google Scholar] [CrossRef]
  32. Faulkner, G.E.; Buliung, R.N.; Flora, P.K.; Fusco, C. Active school transport, physical activity levels and body weight of children and youth: A systematic review. Prev. Med. 2009, 48, 3–8. [Google Scholar] [CrossRef]
  33. Schoeppe, S.; Duncan, M.J.; Badland, H.; Oliver, M.; Curtis, C. Associations of children’s independent mobility and active travel with physical activity, sedentary behaviour and weight status: A systematic review. J. Sci. Med. Sport 2013, 16, 312–319. [Google Scholar] [CrossRef] [PubMed]
  34. Cooper, A.R.; Andersen, L.B.; Wedderkopp, N.; Page, A.S.; Froberg, K. Physical Activity Levels of Children Who Walk, Cycle, or Are Driven to School. Am. J. Prev. Med. 2005, 29, 179–184. [Google Scholar] [CrossRef] [PubMed]
  35. Sirard, J.R.; Riner, W.F., Jr.; McIver, K.L.; Pate, R.R. Physical Activity and Active Commuting to Elementary School. Med. Sci. Sports Exerc. 2005, 37, 2062–2069. [Google Scholar] [CrossRef] [PubMed]
  36. Goodman, A.; Mackett, R.L.; Paskins, J. Activity compensation and activity synergy in British 8–13year olds. Prev. Med. 2011, 53, 293–298. [Google Scholar] [CrossRef] [PubMed]
  37. Mendoza, J.A.; Watson, K.; Nguyen, N.; Cerin, E.; Baranowski, T.; Nicklas, T.A. Active Commuting to School and Association with Physical Activity and Adiposity among US Youth. J. Phys. Act. Health 2011, 8, 488–495. [Google Scholar] [CrossRef] [PubMed]
  38. Denstel, K.D.; Broyles, S.T.; Larouche, R.; Sarmiento, O.L.; Barreira, T.V.; Chaput, J.-P.; Church, T.S.; Fogelholm, M.; Hu, G.; Kuriyan, R.; et al. Active school transport and weekday physical activity in 9–11-year-old children from 12 countries. Int. J. Obes. Suppl. 2015, 5, S100–S106. [Google Scholar] [CrossRef] [PubMed]
  39. Saksvig, B.I.; Catellier, D.J.; Pfeiffer, K.; Schmitz, K.H.; Conway, T.; Going, S.; Ward, D.; Strikmiller, P.; Treuth, M.S. Travel by Walking before and after School and Physical Activity among Adolescent Girls. Arch. Pediatr. Adolesc. Med. 2007, 161, 153–158. [Google Scholar] [CrossRef] [PubMed]
  40. Van Sluijs, E.M.F.; Fearne, V.A.; Mattocks, C.; Riddoch, C.; Griffin, S.J.; Ness, A. The contribution of active travel to children’s physical activity levels: Cross-sectional results from the ALSPAC study. Prev. Med. 2009, 48, 519–524. [Google Scholar] [CrossRef]
  41. Cooper, A.R.; Page, A.S.; Wheeler, B.W.; Griew, P.; Davis, L.; Hillsdon, M.; Jago, R. Mapping the Walk to School Using Accelerometry Combined with a Global Positioning System. Am. J. Prev. Med. 2010, 38, 178–183. [Google Scholar] [CrossRef]
  42. Mendoza, J.A.; Watson, K.; Baranowski, T.; Nicklas, T.A.; Uscanga, D.K.; Nguyen, N.; Hanfling, M.J. Ethnic Minority Children’s Active Commuting to School and Association with Physical Activity and Pedestrian Safety Behaviors. J. Appl. Res. Child 2010, 1, 4. [Google Scholar] [CrossRef]
  43. King, A.C.; Parkinson, K.N.; Adamson, A.J.; Murray, L.; Besson, H.; Reilly, J.J.; Basterfield, L.; the Gateshead Millennium Study Core Team. Correlates of objectively measured physical activity and sedentary behaviour in English children. Eur. J. Public Health 2011, 21, 424–431. [Google Scholar] [CrossRef] [PubMed]
  44. Panter, J.; Jones, A.; Van Sluijs, E.; Griffin, S. The influence of distance to school on the associations between active commuting and physical activity. Pediatr. Exerc. Sci. 2011, 23, 72–86. [Google Scholar] [CrossRef] [PubMed]
  45. Owen, C.G.; Nightingale, C.M.; Rudnicka, A.R.; van Sluijs, E.M.F.; Ekelund, U.; Cook, D.G.; Whincup, P.H. Travel to School and Physical Activity Levels in 9–10 Year-Old UK Children of Different Ethnic Origin; Child Heart and Health Study in England (CHASE). PLoS ONE 2012, 7, e30932. [Google Scholar] [CrossRef] [PubMed]
  46. Roth, M.A.; Millett, C.J.; Mindell, J.S. The contribution of active travel (walking and cycling) in children to overall physical activity levels: A national cross sectional study. Prev. Med. 2012, 54, 134–139. [Google Scholar] [CrossRef] [PubMed]
  47. Abbott, R.A.; Macdonald, D.; Nambiar, S.; Davies, P.S. The association between walking to school, daily step counts and meeting step targets in 5- to 17-year-old Australian children. Pediatr. Exerc. Sci. 2009, 21, 520–532. [Google Scholar] [CrossRef] [PubMed]
  48. McCormack, G.R.; Giles-Corti, B.; Timperio, A.; Wood, G.; Villanueva, K. A cross-sectional study of the individual, social, and built environmental correlates of pedometer-based physical activity among elementary school children. Int. J. Behav. Nutr. Phys. Act. 2011, 8, 30. [Google Scholar] [CrossRef] [PubMed]
  49. Loucaides, C.A.; Jago, R. Differences in physical activity by gender, weight status and travel mode to school in Cypriot children. Prev. Med. 2008, 47, 107–111. [Google Scholar] [CrossRef]
  50. Duncan, E.K.; Duncan, J.S.; Schofield, G. Pedometer-determined physical activity and active transport in girls. Int. J. Behav. Nutr. Phys. Act. 2008, 5, 2. [Google Scholar] [CrossRef]
  51. Hohepa, M.; Schofield, G.; Kolt, G.S.; Scragg, R.; Garrett, N. Pedometer-Determined Physical Activity Levels of Adolescents: Differences by Age, Sex, Time of Week, and Transportation Mode to School. J. Phys. Act. Health 2008, 5 (Suppl. S1), S140–S152. [Google Scholar] [CrossRef]
  52. Larouche, R.; Lloyd, M.; Knight, E.; Tremblay, M.S. Relationship Between Active School Transport and Body Mass Index in Grades-4-to-6 Children. Pediatr. Exerc. Sci. 2011, 23, 322–330. [Google Scholar] [CrossRef]
  53. Johnson, T.G.; Brusseau, T.A.; Darst, P.W.; Kulinna, P.H.; White-Taylor, J. Step Counts of Non-White Minority Children and Youth by Gender, Grade Level, Race/Ethnicity, and Mode of School Transportation. J. Phys. Act. Health 2010, 7, 730–736. [Google Scholar] [CrossRef] [PubMed]
  54. Cuddihy, T.; Davidson, M.; Michaud-Tomson, L. Walk to School—Does it Make a Difference in Children’s Physical Activity Levels? Healthy Lifestyles J. 2003, 50, 16–24. [Google Scholar]
  55. Murtagh, E.M.; Murphy, M.H. Active travel to school and physical activity levels of Irish primary schoolchildren. Pediatr. Exerc. Sci. 2011, 23, 230–236. [Google Scholar] [CrossRef] [PubMed]
  56. Tudor-Locke, C.; Neff, L.J.; Ainsworth, B.E.; Addy, C.L.; Popkin, B.M. Omission of active commuting to school and the prevalence of children’s health-related physical activity levels: The Russian Longitudinal Monitoring Study. Child Care Health Dev. 2002, 28, 507–512. [Google Scholar] [CrossRef] [PubMed]
  57. Heelan, K.A.; Donnelly, J.E.; Jacobsen, D.J.; Mayo, M.S.; Washburn, R.; Greene, L. Active commuting to and from school and BMI in elementary school children—Preliminary data. Child Care Health Dev. 2005, 31, 341–349. [Google Scholar] [CrossRef] [PubMed]
  58. Trang, N.H.H.D.; Hong, T.K.; Dibley, M.J.; Sibbritt, D.W. Factors Associated with Physical Inactivity in Adolescents in Ho Chi Minh City, Vietnam. Med. Sci. Sports Exerc. 2009, 41, 1374–1383. [Google Scholar] [CrossRef] [PubMed]
  59. Daly-Smith, A.J.; McKenna, J.; Radley, D.; Long, J. The impact of additional weekdays of active commuting to school on children achieving a criterion of 300+ minutes of moderate-to-vigorous physical activity. Health Educ. J. 2010, 70, 428–434. [Google Scholar] [CrossRef]
  60. Rosenberg, D.E.; Sallis, J.F.; Conway, T.L.; Cain, K.L.; McKenzie, T.L. Active Transportation to School Over 2 Years in Relation to Weight Status and Physical Activity. Obesity 2006, 14, 1771–1776. [Google Scholar] [CrossRef]
  61. Carver, A.; Timperio, A.F.; Hesketh, K.D.; Ridgers, N.D.; Salmon, J.L.; Crawford, D.A. How is active transport associated with children’s and adolescents’ physical activity over time? Int. J. Behav. Nutr. Phys. Act. 2011, 8, 126. [Google Scholar] [CrossRef]
  62. Andersen, L.B.; Wedderkopp, N.; Kristensen, P.; Moller, N.C.; Froberg, K.; Cooper, A.R. Cycling to School and Cardiovascular Risk Factors: A Longitudinal Study. J. Phys. Act. Health 2011, 8, 1025–1033. [Google Scholar] [CrossRef]
  63. Camiletti-Moirón, D.; Timperio, A.; Veitch, J.; Fernández-Santos, J.d.R.; Abbott, G.; Delgado-Alfonso, A.; Cabanas-Sanchez, V.; Veiga, O.L.; Salmon, J.; Castro-Piñero, J. Changes in and the mediating role of physical activity in relation to active school transport, fitness and adiposity among Spanish youth: The UP&DOWN longitudinal study. Int. J. Behav. Nutr. Phys. Act. 2020, 17, 37. [Google Scholar] [CrossRef] [PubMed]
  64. Alwis, G.; Linden, C.; Dencker, M.; Stenevi-Lundgren, S.; Gardsell, P.; Karlsson, M.K. Bone mineral accrual and gain in skeletal width in pre-pubertal school children is independent of the mode of school transportation—One-year data from the prospective observational pediatric osteoporosis prevention (POP) study. BMC Musculoskelet. Disord. 2007, 8, 66. [Google Scholar] [CrossRef]
  65. Löfgren, B.; Stenevi-Lundgren, S.; Dencker, M.; Karlsson, M.K. The mode of school transportation in pre-pubertal children does not influence the accrual of bone mineral or the gain in bone size—Two year prospective data from the paediatric osteoporosis preventive (POP) study. BMC Musculoskelet. Disord. 2010, 11, 25. [Google Scholar] [CrossRef] [PubMed]
  66. Smith, L.; Sahlqvist, S.; Ogilvie, D.; Jones, A.; Corder, K.; Griffin, S.J.; van Sluijs, E. Is a change in mode of travel to school associated with a change in overall physical activity levels in children? Longitudinal results from the SPEEDY study. Int. J. Behav. Nutr. Phys. Act. 2012, 9, 134. [Google Scholar] [CrossRef] [PubMed]
  67. Heelan, K.A.; Abbey, B.M.; Donnelly, J.E.; Mayo, M.S.; Welk, G.J. Evaluation of a Walking School Bus for Promoting Physical Activity in Youth. J. Phys. Act. Health 2009, 6, 560–567. [Google Scholar] [CrossRef]
  68. Sirard, J.R.; Alhassan, S.; Spencer, T.R.; Robinson, T.N. Changes in Physical Activity from Walking to School. J. Nutr. Educ. Behav. 2008, 40, 324–326. [Google Scholar] [CrossRef] [PubMed]
  69. Mendoza, J.A.; Watson, K.; Baranowski, T.; Nicklas, T.A.; Uscanga, D.K.; Hanfling, M.J. The Walking School Bus and Children’s Physical Activity: A Pilot Cluster Randomized Controlled Trial. Pediatrics 2011, 128, e537–e544. [Google Scholar] [CrossRef] [PubMed]
  70. Cooper, A.R.; Wedderkopp, N.; Wang, H.; Andersen, L.B.; Froberg, K.; Page, A.S. Active Travel to School and Cardiovascular Fitness in Danish Children and Adolescents. Med. Sci. Sports Exerc. 2006, 38, 1724–1731. [Google Scholar] [CrossRef]
  71. Voss, C.; Sandercock, G. Aerobic Fitness and Mode of Travel to School in English Schoolchildren. Med. Sci. Sports Exerc. 2010, 42, 281–287. [Google Scholar] [CrossRef]
  72. Cooper, A.R.; Page, A.S.; Foster, L.J.; Qahwaji, D. Commuting to school: Are children who walk more physically active? Am. J. Prev. Med. 2003, 25, 273–276. [Google Scholar] [CrossRef]
  73. Metcalf, B.; Voss, L.; Jeffery, A.; Perkins, J.; Wilkin, T. Physical activity cost of the school run: Impact on schoolchildren of being driven to school (EarlyBird 22). BMJ 2004, 329, 832–833. [Google Scholar] [CrossRef] [PubMed]
  74. Dollman, J.; Lewis, N.R. Active transport to school as part of a broader habit of walking and cycling among South Australian youth. Pediatr. Exerc. Sci. 2007, 19, 436–443. [Google Scholar] [CrossRef] [PubMed]
  75. Ford, P.; Bailey, R.; Coleman, D.; Woolf-May, K.; Swaine, I. Activity Levels, Dietary Energy Intake, and Body Composition in Children Who Walk to School. Pediatr. Exerc. Sci. 2007, 19, 393–407. [Google Scholar] [CrossRef]
  76. Baig, F.; Hameed, M.; Li, M.; Shorthouse, G.; Roalfe, A.; Daley, A. Association between active commuting to school, weight and physical activity status in ethnically diverse adolescents predominately living in deprived communities. Public Health 2009, 123, 39–41. [Google Scholar] [CrossRef]
  77. Wen, L.M.; Merom, D.; Rissel, C.; Simpson, J.M. Weight status, modes of travel to school and screen time: A cross-sectional survey of children aged 10-13 years in Sydney. Health Promot. J. Aust. 2010, 21, 57–63. [Google Scholar] [CrossRef]
  78. Masoumi, H. Active Transport to School and Children’s Body Weight: A Systematic Reivew. TeMA-J. Land Use Mobil. Environ. 2017, 10, 95–110. [Google Scholar] [CrossRef]
  79. Xu, H.; Wen, L.M.; Rissel, C. The Relationships Between Active Transport to Work or School and Cardiovascular Health or Body Weight: A systematic review. Asia Pac. J. Public Health 2013, 25, 298–315. [Google Scholar] [CrossRef] [PubMed]
  80. Lubans, D.R.; Boreham, C.A.; Kelly, P.; Foster, C.E. The relationship between active travel to school and health-related fitness in children and adolescents: A systematic review. Int. J. Behav. Nutr. Phys. Act. 2011, 8, 5. [Google Scholar] [CrossRef]
  81. Ruiz-Hermosa, A.; Martínez-Vizcaíno, V.; Alvarez-Bueno, C.; García-Prieto, J.C.; Pardo-Guijarro, M.J.; Sánchez-López, M. No Association Between Active Commuting to School, Adiposity, Fitness, and Cognition in Spanish Children: The MOVI-KIDS Study. J. Sch. Health 2018, 88, 839–846. [Google Scholar] [CrossRef]
  82. Mwaikambo, S.A.; Leyna, G.H.; Killewo, J.; Simba, A.; Puoane, T. Why are primary school children overweight and obese? A cross sectional study undertaken in Kinondoni district, Dar-es-salaam. BMC Public Health 2015, 15, 1269. [Google Scholar] [CrossRef]
  83. Manyanga, T.; Barnes, J.D.; Chaput, J.-P.; Dubois, L.; Katzmarzyk, P.T.; Mire, E.F.; Prista, A.; Tremblay, M.S. Prevalence and correlates of objectively measured weight status among urban and rural Mozambican primary schoolchildren: A cross-sectional study. PLoS ONE 2020, 15, e0228592. [Google Scholar] [CrossRef] [PubMed]
  84. Silva, K.S.; Lopes, A.S. Excess weight, arterial pressure and physical activity in commuting to school: Correlations. Arq. Bras. Cardiol. 2008, 91, 84–91. [Google Scholar] [PubMed]
  85. Panter, J.R.; Jones, A.P.; van Sluijs, E.M.F.; Griffin, S.J. Attitudes, social support and environmental perceptions as predictors of active commuting behaviour in school children. J. Epidemiol. Community Health 2009, 64, 41–48. [Google Scholar] [CrossRef] [PubMed]
  86. Andegiorgish, A.K.; Wang, J.; Zhang, X.; Liu, X.; Zhu, H. Prevalence of overweight, obesity, and associated risk factors among school children and adolescents in Tianjin, China. Eur. J. Pediatr. 2012, 171, 697–703. [Google Scholar] [CrossRef] [PubMed]
  87. Ostergaard, L.; Kolle, E.; Steene-Johannessen, J.; Anderssen, S.A.; Andersen, L.B. Cross sectional analysis of the association between mode of school transportation and physical fitness in children and adolescents. Int. J. Behav. Nutr. Phys. Act. 2013, 10, 91. [Google Scholar] [CrossRef] [PubMed]
  88. Silva, K.S.; Vasques, D.G.; Martins, C.d.O.; Williams, L.A.; Lopes, A.S. Active Commuting: Prevalence, Barriers, and Associated Variables. J. Phys. Act. Health 2011, 8, 750–757. [Google Scholar] [CrossRef] [PubMed]
  89. Noonan, R.J.; Boddy, L.M.; Knowles, Z.R.; Fairclough, S.J. Fitness, Fatness and Active School Commuting among Liverpool Schoolchildren. Int. J. Environ. Res. Public Health 2017, 14, 995. [Google Scholar] [CrossRef]
  90. Sun, Y.; Liu, Y.; Tao, F.-B. Associations Between Active Commuting to School, Body Fat, and Mental Well-being: Population-Based, Cross-Sectional Study in China. J. Adolesc. Health 2015, 57, 679–685. [Google Scholar] [CrossRef]
  91. Klein-Platat, C.; Oujaa, M.; Wagner, A.; Haan, M.C.; Arveiler, D.; Schlienger, J.L.; Simon, C. Physical activity is inversely related to waist circumference in 12-y-old French adolescents. Int. J. Obes. 2005, 29, 9–14. [Google Scholar] [CrossRef]
  92. Smith, L.; Aggio, D.; Hamer, M. Active travel to non-school destinations but not to school is associated with higher physical activity levels in an ethnically diverse sample of inner-city schoolchildren. BMC Public Health 2017, 17, 13. [Google Scholar] [CrossRef]
  93. Cooper, A.R.; Wedderkopp, N.; Jago, R.; Kristensen, P.L.; Moller, N.C.; Froberg, K.; Page, A.S.; Andersen, L.B. Longitudinal associations of cycling to school with adolescent fitness. Prev. Med. 2008, 47, 324–328. [Google Scholar] [CrossRef] [PubMed]
  94. Chillón, P.; Ortega, F.B.; Ruiz, J.R.; Evenson, K.R.; Labayen, I.; Martínez-Vizcaino, V.; Hurtig-Wennlöf, A.; Veidebaum, T.; Sjöström, M. Bicycling to school is associated with improvements in physical fitness over a 6-year follow-up period in Swedish children. Prev. Med. 2012, 55, 108–112. [Google Scholar] [CrossRef] [PubMed]
  95. Chillón, P.; Ortega, F.B.; Ruiz, J.R.; Veidebaum, T.; Oja, L.; Mäestu, J.; Sjöström, M. Active commuting to school in children and adolescents: An opportunity to increase physical activity and fitness. Scand. J. Public Health 2010, 38, 873–879. [Google Scholar] [CrossRef] [PubMed]
  96. Børrestad, L.A.; Østergaard, L.; Andersen, L.B.; Bere, E. Experiences from a randomised, controlled trial on cycling to school: Does cycling increase cardiorespiratory fitness? Scand. J. Public Health 2012, 40, 245–252. [Google Scholar] [CrossRef] [PubMed]
  97. Muntaner-Mas, A.; Herrador-Colmenero, M.; Borràs, P.A.; Chillón, P. Physical activity, but not active commuting to school, is associated with cardiorespiratory fitness levels in young people. J. Transp. Health 2018, 10, 297–303. [Google Scholar] [CrossRef]
  98. Meron, D.; Rissel, C.; Reinten-Reynolds, T.; Hardy, L.L. Changes in active travel of school children from 2004 to 2010 in New South Wales, Australia. Prev. Med. 2011, 53, 408–410. [Google Scholar] [CrossRef] [PubMed]
  99. Aires, L.; Pratt, M.; Lobelo, F.; Santos, R.M.; Santos, M.P.; Mota, J. Associations of cardiorespiratory fitness in children and adolescents with physical activity, active commuting to school, and screen time. J. Phys. Act. Health 2011, 8 (Suppl. S2), S198–S205. [Google Scholar] [CrossRef]
  100. Sandercock, G.R.; Ogunleye, A.A. Screen time and passive school travel as independent predictors of cardiorespiratory fitness in youth. Prev. Med. 2012, 54, 319–322. [Google Scholar] [CrossRef]
  101. Villa-González, E.; Ruiz, J.R.; Chillón, P. Associations between Active Commuting to School and Health-Related Physical Fitness in Spanish School-Aged Children: A Cross-Sectional Study. Int. J. Environ. Res. Public Health 2015, 12, 10362–10373. [Google Scholar] [CrossRef]
  102. Cohen, D.; Ogunleye, A.A.; Taylor, M.; Voss, C.; Micklewright, D.; Sandercock, G.R. Association between habitual school travel and muscular fitness in youth. Prev. Med. 2014, 67, 216–220. [Google Scholar] [CrossRef]
  103. Pires, A.L.; Soares, S.S.; Welser, L.; da Silva, C.F.; Sehn, A.P.; Kern, D.G.; Valim, A.R.D.; Reuter, C.P.; Burgos, M.S. Associação entre o deslocamento para a escola e aptidão física relacionada ao desempenho motor em escolares. Rev. Epidemiol. Controle Infecção 2017, 7, 189–193. [Google Scholar] [CrossRef]
  104. Henriques-Neto, D.; Peralta, M.; Garradas, S.; Pelegrini, A.; Pinto, A.A.; Sánchez-Miguel, P.A.; Marques, A. Active Commuting and Physical Fitness: A Systematic Review. Int. J. Environ. Res. Public Health 2020, 17, 2721. [Google Scholar] [CrossRef] [PubMed]
  105. López-Vicente, M.; Forns, J.; Esnaola, M.; Suades-González, E.; Álvarez-Pedrerol, M.; Robinson, O.; Júlvez, J.; Garcia-Aymerich, J.; Sunyer, J. Physical Activity and Cognitive Trajectories in Schoolchildren. Pediatr. Exerc. Sci. 2016, 28, 431–438. [Google Scholar] [CrossRef] [PubMed]
  106. Westman, J.; Olsson, L.E.; Gärling, T.; Friman, M. Children’s travel to school: Satisfaction, current mood, and cognitive performance. Transportation 2016, 44, 1365–1382. [Google Scholar] [CrossRef]
  107. Fang, J.-T.; Lin, J.-J. School travel modes and children’s spatial cognition. Urban Stud. 2016, 54, 1578–1600. [Google Scholar] [CrossRef]
  108. Moran, M.R.; Eizenberg, E.; Plaut, P. Getting to Know a Place: Built Environment Walkability and Children’s Spatial Representation of Their Home-School (h–s) Route. Int. J. Environ. Res. Public Health 2017, 14, 607. [Google Scholar] [CrossRef] [PubMed]
  109. Rissotto, A.; Tonucci, F. Freedom of movement and environmental knowledge in elementary school children. J. Environ. Psychol. 2002, 22, 65–77. [Google Scholar] [CrossRef]
  110. Haapala, E.A.; Poikkeus, A.-M.; Kukkonen-Harjula, K.; Tompuri, T.; Lintu, N.; Väistö, J.; Leppänen, P.H.T.; Laaksonen, D.E.; Lindi, V.; Lakka, T.A. Associations of Physical Activity and Sedentary Behavior with Academic Skills—A Follow-Up Study among Primary School Children. PLoS ONE 2014, 9, e107031. [Google Scholar] [CrossRef]
  111. Domazet, S.L.; Tarp, J.; Huang, T.; Gejl, A.K.; Andersen, L.B.; Froberg, K.; Bugge, A. Associations of Physical Activity, Sports Participation and Active Commuting on Mathematic Performance and Inhibitory Control in Adolescents. PLoS ONE 2016, 11, e0146319. [Google Scholar] [CrossRef]
  112. García-Hermoso, A.; Saavedra, J.M.; Olloquequi, J.; Ramírez-Vélez, R. Associations between the duration of active commuting to school and academic achievement in rural Chilean adolescents. Environ. Health Prev. Med. 2017, 22, 31. [Google Scholar] [CrossRef]
  113. Mora-Gonzalez, J.; Rodríguez-López, C.; Cadenas-Sanchez, C.; Herrador-Colmenero, M.; Esteban-Cornejo, I.; Huertas-Delgado, F.J.; Ardoy, D.N.; Ortega, F.B.; Chillón, P. Active commuting to school was inversely associated with academic achievement in primary but not secondary school students. Acta Paediatr. 2017, 106, 334–340. [Google Scholar] [CrossRef] [PubMed]
  114. Stark, J.; Meschik, M.; Singleton, P.A.; Schützhofer, B. Active school travel, attitudes and psychological well-being of children. Transp. Res. Part F Traffic Psychol. Behav. 2018, 56, 453–465. [Google Scholar] [CrossRef]
  115. Pizarro, A.N.; Ribeiro, J.C.; Marques, E.A.; Mota, J.; Santos, M.P. Is walking to school associated with improved metabolic health? Int. J. Behav. Nutr. Phys. Act. 2013, 10, 12. [Google Scholar] [CrossRef] [PubMed]
  116. Gutiérrez-Zornoza, M.; Sánchez-López, M.; García-Hermoso, A.; González-García, A.; Chillón, P.; Martínez-Vizcaíno, V. Active Commuting to School, Weight Status, and Cardiometabolic Risk in Children from Rural Areas: The Cuenca study. Health Educ. Behav. 2015, 42, 231–239. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, X.; Smith, N.A.; Sumowski, M.T.; Anderson, J.M.; Anderson, K.; Badenoch, E.A.; Brady, S.J.; Coleman, M.; Coull, R.F.; Green, D.; et al. Active travelling to school is not associated with increased total daily physical activity levels, or reduced obesity and cardiovascular/pulmonary health parameters in 10–12-year olds: A cross-sectional cohort study. Int. J. Obes. 2020, 44, 1452–1466. [Google Scholar] [CrossRef] [PubMed]
  118. Ginja, S.; Arnott, B.; Namdeo, A.; McColl, E. Understanding active school travel through the Behavioural Ecological Model. Health Psychol. Rev. 2018, 12, 58–74. [Google Scholar] [CrossRef] [PubMed]
  119. Lu, W.; McKyer, E.L.J.; Lee, C.; Goodson, P.; Ory, M.G.; Wang, S. Perceived barriers to children’s active commuting to school: A systematic review of empirical, methodological and theoretical evidence. Int. J. Behav. Nutr. Phys. Act. 2014, 11, 140. [Google Scholar] [CrossRef] [PubMed]
  120. McMillan, T.E. Urban Form and a Child’s Trip to School: The Current Literature and a Framework for Future Research. J. Plan. Lit. 2005, 19, 440–456. [Google Scholar] [CrossRef]
  121. Pont, K.; Ziviani, J.; Wadley, D.; Bennett, S.; Abbott, R. Environmental correlates of children’s active transportation: A systematic literature review. Health Place 2009, 15, 827–840. [Google Scholar] [CrossRef]
  122. D’haese, S.; Vanwolleghem, G.; Hinckson, E.; De Bourdeaudhuij, I.; Deforche, B.; Van Dyck, D.; Cardon, G. Cross-continental comparison of the association between the physical environment and active transportation in children: A systematic review. Int. J. Behav. Nutr. Phys. Act. 2015, 12, 145. [Google Scholar] [CrossRef]
  123. Wangzom, D.; White, M.; Paay, J. Perceived Safety Influencing Active Travel to School—A Built Environment Perspective. Int. J. Environ. Res. Public Health 2023, 20, 1026. [Google Scholar] [CrossRef] [PubMed]
  124. Aranda-Balboa, M.J.; Huertas-Delgado, F.J.; Herrador-Colmenero, M.; Cardon, G.; Chillón, P. Parental barriers to active transport to school: A systematic review. Int. J. Public Health 2020, 65, 87–98. [Google Scholar] [CrossRef] [PubMed]
  125. Rothman, L.; Macpherson, A.K.; Ross, T.; Buliung, R.N. The decline in active school transportation (AST): A systematic review of the factors related to AST and changes in school transport over time in North America. Prev. Med. 2018, 111, 314–322. [Google Scholar] [CrossRef] [PubMed]
  126. Wong, B.Y.-M.; Faulkner, G.; Buliung, R. GIS measured environmental correlates of active school transport: A systematic review of 14 studies. Int. J. Behav. Nutr. Phys. Act. 2011, 8, 39. [Google Scholar] [CrossRef] [PubMed]
  127. Ikeda, E.; Hinckson, E.; Witten, K.; Smith, M. Associations of children’s active school travel with perceptions of the physical environment and characteristics of the social environment: A systematic review. Health Place 2018, 54, 118–131. [Google Scholar] [CrossRef] [PubMed]
  128. Nikitas, A.; Wang, J.Y.; Knamiller, C. Exploring parental perceptions about school travel and walking school buses: A thematic analysis approach. Transp. Res. Part A Policy Pract. 2019, 124, 468–487. [Google Scholar] [CrossRef]
  129. Forsberg, H.; Rutberg, S.; Mikaelsson, K.; Lindqvist, A.-K. It’s about being the good parent: Exploring attitudes and beliefs towards active school transportation. Int. J. Circumpolar Health 2020, 79, 1798113. [Google Scholar] [CrossRef] [PubMed]
  130. O’Connor, J.; Brown, A. A qualitative study of ‘fear’ as a regulator of children’s independent physical activity in the suburbs. Health Place 2013, 24, 157–164. [Google Scholar] [CrossRef]
  131. Ahern, S.M.; Arnott, B.; Chatterton, T.; de Nazelle, A.; Kellar, I.; McEachan, R.R. Understanding parents’ school travel choices: A qualitative study using the Theoretical Domains Framework. J. Transp. Health 2017, 4, 278–293. [Google Scholar] [CrossRef]
  132. Faulkner, G.E.; Richichi, V.; Buliung, R.N.; Fusco, C.; Moola, F. What’s “quickest and easiest?”: Parental decision making about school trip mode. Int. J. Behav. Nutr. Phys. Act. 2010, 7, 62. [Google Scholar] [CrossRef]
  133. Wilson, K.; Clark, A.F.; Gilliland, J.A. Understanding child and parent perceptions of barriers influencing children’s active school travel. BMC Public Health 2018, 18, 1053. [Google Scholar] [CrossRef] [PubMed]
  134. Morris, S.; Lawlor, E.R.; Foley, L.; Summerbell, C.; Panter, J.; Adams, J.; Jago, R.; Pollard, T.M. Children’s experiences of the journey between home and school: A qualitative synthesis using meta-ethnography. Health Place 2022, 76, 102819. [Google Scholar] [CrossRef] [PubMed]
  135. Zhu, X.; Lee, C. Correlates of Walking to School and Implications for Public Policies: Survey Results from Parents of Elementary School Children in Austin, Texas. J. Public Health Policy 2009, 30 (Suppl. S1), S177–S202. [Google Scholar] [CrossRef] [PubMed]
  136. Zhu, X.; Arch, B.; Lee, C. Personal, Social, and Environmental Correlates of Walking to School Behaviors: Case Study in Austin, Texas. Sci. World J. 2008, 8, 859–872. [Google Scholar] [CrossRef] [PubMed]
  137. Gustat, J.; Richards, K.; Rice, J.; Andersen, L.; Parker-Karst, K.; Cole, S. Youth Walking and Biking Rates Vary by Environments Around 5 Louisiana Schools. J. Sch. Health 2015, 85, 36–42. [Google Scholar] [CrossRef] [PubMed]
  138. Ahlport, K.N.; Linnan, L.; Vaughn, A.; Evenson, K.R.; Ward, D.S. Barriers to and Facilitators of Walking and Bicycling to School: Formative Results from the Non-Motorized Travel Study. Health Educ. Behav. 2008, 35, 221–244. [Google Scholar] [CrossRef] [PubMed]
  139. Rothman, L.; Buliung, R.; To, T.; Macarthur, C.; Macpherson, A.; Howard, A. Associations between parents’ perception of traffic danger, the built environment and walking to school. J. Transp. Health 2015, 2, 327–335. [Google Scholar] [CrossRef]
  140. Marzi, I.; Emmerling, S.; Demetriou, Y.; Bucksch, J.; Schulze, C.; Brindley, C.; Reimers, A.K. Interventions Aiming to Promote Active Commuting in Children and Adolescents: An Evaluation from a Sex/Gender Perspective. Front. Sports Act. Living 2020, 2, 590857. [Google Scholar] [CrossRef]
  141. Sirard, J.R.; Slater, M.E. Walking and Bicycling to School: A Review. Am. J. Lifestyle Med. 2008, 2, 372–396. [Google Scholar] [CrossRef]
  142. Zuniga, K.D. From barrier elimination to barrier negotiation: A qualitative study of parents’ attitudes about active travel for elementary school trips. Transp. Policy 2012, 20, 75–81. [Google Scholar] [CrossRef]
  143. Safe Routes Partnership. Dropping Enforcement from the Safe Routes to School 6 E’s Framework. Available online: https://www.saferoutespartnership.org/blog/dropping-enforcement-safe-routes-school-6-e%E2%80%99s-framework (accessed on 6 April 2023).
  144. Smith, L.; Norgate, S.H.; Cherrett, T.; Davies, N.; Winstanley, C.; Harding, M. Walking School Buses as a Form of Active Transportation for Children—A Review of the Evidence. J. Sch. Health 2015, 85, 197–210. [Google Scholar] [CrossRef] [PubMed]
  145. Buliung, R.; Faulkner, G.; Beesley, T.; Kennedy, J. School Travel Planning: Mobilizing School and Community Resources to Encourage Active School Transportation. J. Sch. Health 2011, 81, 704–712. [Google Scholar] [CrossRef]
  146. Chillón, P.; Evenson, K.R.; Vaughn, A.; Ward, D.S. A systematic review of interventions for promoting active transportation to school. Int. J. Behav. Nutr. Phys. Act. 2011, 8, 10. [Google Scholar] [CrossRef] [PubMed]
  147. Pang, B.; Kubacki, K.; Rundle-Thiele, S. Promoting active travel to school: A systematic review (2010–2016). BMC Public Health 2017, 17, 638. [Google Scholar] [CrossRef] [PubMed]
  148. Larouche, R.; Mammen, G.; Rowe, D.A.; Faulkner, G. Effectiveness of active school transport interventions: A systematic review and update. BMC Public Health 2018, 18, 206. [Google Scholar] [CrossRef] [PubMed]
  149. Villa-González, E.; Barranco-Ruiz, Y.; Evenson, K.R.; Chillón, P. Systematic review of interventions for promoting active school transport. Prev. Med. 2018, 111, 115–134. [Google Scholar] [CrossRef] [PubMed]
  150. Jones, R.A.; Blackburn, N.E.; Woods, C.; Byrne, M.; van Nassau, F.; Tully, M.A. Interventions promoting active transport to school in children: A systematic review and meta-analysis. Prev. Med. 2019, 123, 232–241. [Google Scholar] [CrossRef]
  151. Schönbach, D.M.I.; Altenburg, T.M.; Marques, A.; Chinapaw, M.J.M.; Demetriou, Y. Strategies and effects of school-based interventions to promote active school transportation by bicycle among children and adolescents: A systematic review. Int. J. Behav. Nutr. Phys. Act. 2020, 17, 138. [Google Scholar] [CrossRef]
  152. Jacob, V.; Chattopadhyay, S.K.; Reynolds, J.A.; Hopkins, D.P.; Morgan, J.A.; Brown, D.R.; Kochtitzky, C.S.; Cuellar, A.E.; Kumanyika, S.K.; Community Preventive Services Task Force. Economics of Interventions to Increase Active Travel to School: A Community Guide Systematic Review. Am. J. Prev. Med. 2021, 60, e27–e40. [Google Scholar] [CrossRef]
  153. Jacobs, J.; Crooks, N.; Allender, S.; Strugnell, C.; Backholer, K.; Nichols, M. Is the physical activity environment surrounding primary schools associated with students’ weight status, physical activity or active transport, in regional areas of Victoria, Australia? A cross-sectional study. BMJ Open 2021, 11, e045785. [Google Scholar] [CrossRef]
  154. Merom, D.; Tudor-Locke, C.; Bauman, A.; Rissel, C. Active commuting to school among NSW primary school children: Implications for public health. Health Place 2006, 12, 678–687. [Google Scholar] [CrossRef] [PubMed]
  155. Bringolf-Isler, B.; Grize, L.; Mäder, U.; Ruch, N.; Sennhauser, F.H.; Braun-Fahrländer, C.; SCARPOL Team. Personal and environmental factors associated with active commuting to school in Switzerland. Prev. Med. 2008, 46, 67–73. [Google Scholar] [CrossRef] [PubMed]
  156. Macdonald, L.; McCrorie, P.; Nicholls, N.; Olsen, J.R. Active commute to school: Does distance from school or walkability of the home neighbourhood matter? A national cross-sectional study of children aged 10–11 years, Scotland, UK. BMJ Open 2019, 9, e033628. [Google Scholar] [CrossRef] [PubMed]
  157. Ikeda, E.; Hinckson, E.; Witten, K.; Smith, M. Assessment of direct and indirect associations between children active school travel and environmental, household and child factors using structural equation modelling. Int. J. Behav. Nutr. Phys. Act. 2019, 16, 32. [Google Scholar] [CrossRef] [PubMed]
  158. Dessing, D.; de Vries, S.I.; Graham, J.M.; Pierik, F.H. Active transport between home and school assessed with GPS: A cross-sectional study among Dutch elementary school children. BMC Public Health 2014, 14, 227. [Google Scholar] [CrossRef] [PubMed]
  159. Villa-González, E.; Rosado-López, S.; Barranco-Ruiz, Y.; Herrador-Colmenero, M.; Cadenas-Sánchez, C.; Santos, M.P.; Chillón, P. Objective Measurement of the Mode of Commuting to School Using GPS: A Pilot Study. Sustainability 2019, 11, 5395. [Google Scholar] [CrossRef]
  160. Saelens, B.E.; Sallis, J.F.; Black, J.B.; Chen, D. Neighborhood-Based Differences in Physical Activity: An Environment Scale Evaluation. Am. J. Public Health 2003, 93, 1552–1558. [Google Scholar] [CrossRef]
  161. Cerin, E.; Saelens, B.E.; Sallis, J.F.; Frank, L.D. Neighborhood Environment Walkability Scale: Validity and Development of a Short Form. Med. Sci. Sports Exerc. 2006, 38, 1682–1691. [Google Scholar] [CrossRef]
  162. Rosenberg, D.; Ding, D.; Sallis, J.F.; Kerr, J.; Norman, G.J.; Durant, N.; Harris, S.K.; Saelens, B.E. Neighborhood Environment Walkability Scale for Youth (NEWS-Y): Reliability and relationship with physical activity. Prev. Med. 2009, 49, 213–218. [Google Scholar] [CrossRef]
  163. Mah, S.; Nettlefold, L.; Macdonald, H.; Winters, M.; Race, D.; Voss, C.; McKay, H. Does parental support influence children’s active school travel? Prev. Med. Rep. 2017, 6, 346–351. [Google Scholar] [CrossRef]
  164. Ozbil, A.; Yesiltepe, D.; Argin, G.; Rybarczyk, G. Children’s Active School Travel: Examining the Combined Perceived and Objective Built-Environment Factors from Space Syntax. Int. J. Environ. Res. Public Health 2021, 18, 286. [Google Scholar] [CrossRef] [PubMed]
  165. Larouche, R.; Sarmiento, O.L.; Broyles, S.T.; Denstel, K.D.; Church, T.S.; Barreira, T.V.; Chaput, J.-P.; Fogelholm, M.; Hu, G.; Kuriyan, R.; et al. Are the correlates of active school transport context-specific? Int. J. Obes. Suppl. 2015, 5, S89–S99. [Google Scholar] [CrossRef] [PubMed]
  166. Van Kann, D.H.H.; Kremers, S.P.J.; Gubbels, J.S.; Bartelink, N.H.M.; de Vries, S.I.; de Vries, N.K.; Jansen, M.W.J. The Association Between the Physical Environment of Primary Schools and Active School Transport. Environ. Behav. 2014, 47, 418–435. [Google Scholar] [CrossRef]
  167. Balboa, M.J.A.; Fernández, M.; Villa-González, E.; Murillo-Pardo, B.; Segura-Diaz, J.M.; Saucedo-Araujo, R.G.; Barranco-Ruiz, Y.; Herrador-Colmenero, M.; Huertas-Delgado, F.J.; Chillón, P. O1-4 Validation of questions to assess the mode of commuting to/from school for children according to their parents: The PACO study. Eur. J. Public Health 2022, 32, ckac094.004. [Google Scholar] [CrossRef]
  168. Chillón, P.; Herrador-Colmenero, M.; Migueles, J.H.; Cabanas-Sánchez, V.; Fernandez-Santos, J.R.; Veiga, O.L.; Castro-Piñero, J. Convergent validation of a questionnaire to assess the mode and frequency of commuting to and from school. Scand. J. Public Health 2017, 45, 612–620. [Google Scholar] [CrossRef] [PubMed]
  169. Aranda-Balboa, M.J.; Fernández, M.; Villa-González, E.; Murillo-Pardo, B.; Segura-Díaz, J.M.; Saucedo-Araujo, R.G.; Barranco-Ruiz, Y.; Herrador-Colmenero, M.; Huertas-Delgado, F.J.; Chillón, P. Psychometric Characteristics of a Commuting-to-School Behaviour Questionnaire for Families. Int. J. Environ. Res. Public Health 2020, 17, 8584. [Google Scholar] [CrossRef] [PubMed]
  170. Easton, S.; Ferrari, E. Children’s travel to school—The interaction of individual, neighbourhood and school factors. Transp. Policy 2015, 44, 9–18. [Google Scholar] [CrossRef]
  171. Bosch, L.; Wells, J.C.K.; Lum, S.; Reid, A.M. Associations of the objective built environment along the route to school with children’s modes of commuting: A multilevel modelling analysis (the SLIC study). PLoS ONE 2020, 15, e0231478. [Google Scholar] [CrossRef]
  172. Aarts, M.-J.; Mathijssen, J.J.P.; van Oers, J.A.M.; Schuit, A.J. Associations Between Environmental Characteristics and Active Commuting to School Among Children: A Cross-sectional Study. Int. J. Behav. Med. 2012, 20, 538–555. [Google Scholar] [CrossRef]
  173. Trapp, G.S.A.; Giles-Corti, B.; Christian, H.E.; Bulsara, M.; Timperio, A.F.; McCormack, G.R.; Villaneuva, K.P. Increasing children’s physical activity: Individual, social, and environmental factors associated with walking to and from school. Health Educ. Behav. 2012, 39, 172–182. [Google Scholar] [CrossRef]
  174. Moodie, M.; Haby, M.; Galvin, L.; Swinburn, B.; Carter, R. Cost-effectiveness of active transport for primary school children—Walking School Bus program. Int. J. Behav. Nutr. Phys. Act. 2009, 6, 63. [Google Scholar] [CrossRef]
  175. Tracy, M.; Cerdá, M.; Keyes, K.M. Agent-Based Modeling in Public Health: Current Applications and Future Directions. Annu. Rev. Public Health 2018, 39, 77–94. [Google Scholar] [CrossRef] [PubMed]
  176. Yang, Y.; Diez-Roux, A.V. Using an agent-based model to simulate children’s active travel to school. Int. J. Behav. Nutr. Phys. Act. 2013, 10, 67. [Google Scholar] [CrossRef] [PubMed]
  177. Yang, Y.; Diez-Roux, A.; Evenson, K.R.; Colabianchi, N. Examining the Impact of the Walking School Bus with an Agent-Based Model. Am. J. Public Health 2014, 104, 1196–1203. [Google Scholar] [CrossRef] [PubMed]
  178. Jing, P.; You, Q.; Chen, L. Agent-Based Simulation of Children’s School Travel Mode with Parental Escort Decisions. Information 2018, 9, 50. [Google Scholar] [CrossRef]
  179. Almagor, J.; Martin, A.; McCrorie, P.; Mitchell, R. How can an agent-based model explore the impact of interventions on children’s physical activity in an urban environment? Health Place 2021, 72, 102688. [Google Scholar] [CrossRef] [PubMed]
  180. Herrador-Colmenero, M.; Pérez-García, M.; Ruiz, J.R.; Chillón, P. Assessing Modes and Frequency of Commuting to School in Youngsters: A Systematic Review. Pediatr. Exerc. Sci. 2014, 26, 291–341. [Google Scholar] [CrossRef] [PubMed]
  181. Bushman, B.; Wang, M.C. Vote-counting procedures in meta-analysis. In The Handbook of Research Synthesis and Meta-Analysis; Russell Sage Foundation: New York, NY, USA, 2009; pp. 207–220. [Google Scholar]
  182. Ruiz-Hermosa, A.; Álvarez-Bueno, C.; Cavero-Redondo, I.; Martínez-Vizcaíno, V.; Redondo-Tébar, A.; Sánchez-López, M. Active Commuting to and from School, Cognitive Performance, and Academic Achievement in Children and Adolescents: A Systematic Review and Meta-Analysis of Observational Studies. Int. J. Environ. Res. Public Health 2019, 16, 1839. [Google Scholar] [CrossRef] [PubMed]
  183. Phansikar, M.; Ashrafi, S.A.; Khan, N.A.; Massey, W.V.; Mullen, S.P. Active Commute in Relation to Cognition and Academic Achievement in Children and Adolescents: A Systematic Review and Future Recommendations. Int. J. Environ. Res. Public Health 2019, 16, 5103. [Google Scholar] [CrossRef]
  184. Ikeda, E.; Mavoa, S.; Cavadino, A.; Carroll, P.; Hinckson, E.; Witten, K.; Smith, M. Keeping kids safe for active travel to school: A mixed method examination of school policies and practices and children’s school travel behaviour. Travel Behav. Soc. 2020, 21, 57–68. [Google Scholar] [CrossRef]
  185. Reimers, A.K.; Marzi, I.; Beck, F.; Engels, E.; Renninger, D.; Buttazzoni, A.; Krieger, C.; Demetriou, Y. Active travel behaviour in the family environment: Protocol for the mixed-methods cross-sectional ARRIVE study. BMJ Open 2022, 12, e056383. [Google Scholar] [CrossRef]
  186. Tariq, S.; Woodman, J. Using mixed methods in health research. JRSM Short Rep. 2013, 4, 2042533313479197. [Google Scholar] [CrossRef]
Figure 1. The literature search flowchart.
Figure 1. The literature search flowchart.
Ijerph 20 06929 g001
Figure 2. Global prevalence of ASC among primary school children. Prevalence refers to percentage of school children who commute to or from school using active modes of transportation. Furthermore, additional government data from Australia, Finland, New Zealand, the United Kingdom, and the United States have been included here as they are unavailable in the published literature. Data are included if they cover school children aged between 6 and 12. For surveys which report breakdowns of different modes of transportation, data of different “active” modes are combined to give an overall prevalence of ASC. Caution should be exercised when comparing between countries or regions, given the marked heterogeneity in how ASC is defined and measured, as well as the variations in target age groups across different surveys (further discussed in Section 3.7). Source of data (in alphabetical order): Australia—CensusAtSchool [17]; Australia (New South Wales)—Household Travel Surveys [23]; Brazil (São Paulo)—São Paulo Metropolitan Area Household Travel Survey [24]; Canada (Greater Toronto Area) [25]; China (9 provinces)—China Health and Nutrition Surveys [26]; Czech Republic—Health Behaviour in School-aged Children study [18]; Finland [19,20,21]; New Zealand—The New Zealand Household Travel Survey [13]; Republic of Ireland—The Children’s Sport Participation and Physical Activity Study [28]; Spain—PACO Study [22]; Switzerland—Swiss Microcensus on Travel Behaviour [14]; UK—UK National Travel Survey [15,29]; USA—National Household Travel Survey [16].
Figure 2. Global prevalence of ASC among primary school children. Prevalence refers to percentage of school children who commute to or from school using active modes of transportation. Furthermore, additional government data from Australia, Finland, New Zealand, the United Kingdom, and the United States have been included here as they are unavailable in the published literature. Data are included if they cover school children aged between 6 and 12. For surveys which report breakdowns of different modes of transportation, data of different “active” modes are combined to give an overall prevalence of ASC. Caution should be exercised when comparing between countries or regions, given the marked heterogeneity in how ASC is defined and measured, as well as the variations in target age groups across different surveys (further discussed in Section 3.7). Source of data (in alphabetical order): Australia—CensusAtSchool [17]; Australia (New South Wales)—Household Travel Surveys [23]; Brazil (São Paulo)—São Paulo Metropolitan Area Household Travel Survey [24]; Canada (Greater Toronto Area) [25]; China (9 provinces)—China Health and Nutrition Surveys [26]; Czech Republic—Health Behaviour in School-aged Children study [18]; Finland [19,20,21]; New Zealand—The New Zealand Household Travel Survey [13]; Republic of Ireland—The Children’s Sport Participation and Physical Activity Study [28]; Spain—PACO Study [22]; Switzerland—Swiss Microcensus on Travel Behaviour [14]; UK—UK National Travel Survey [15,29]; USA—National Household Travel Survey [16].
Ijerph 20 06929 g002
Figure 3. McMillan’s Conceptual Framework of an Elementary-Aged Child’s Travel Behaviour (adapted from McMillan [120]).
Figure 3. McMillan’s Conceptual Framework of an Elementary-Aged Child’s Travel Behaviour (adapted from McMillan [120]).
Ijerph 20 06929 g003
Figure 4. Summary of key review findings and future research implications.
Figure 4. Summary of key review findings and future research implications.
Ijerph 20 06929 g004
Table 1. Search terms/strings utilised for identifying pertinent literature.
Table 1. Search terms/strings utilised for identifying pertinent literature.
Electronic DatabasesSearch TermsNumber of Articles
PubMed#1: schoolchild OR child
#2: school OR primary school OR elementary school OR junior school
#3: active transport OR active commute OR active travel OR walk OR bike OR cycle OR bicycle OR skateboard OR vehicle OR motorised transport OR car OR drive
#4: #1 AND #2 AND #3
40,594
Scopus(schoolchild OR child) AND (“school” OR “primary school” OR “elementary school” OR “junior school”) AND (“active transport” OR “active commute” OR “active travel”) AND (walk OR bike OR cycle OR bicycle OR skateboard OR vehicle OR “motorised transport” OR car OR drive)256
ScienceDirect(schoolchild OR child) AND (school OR “primary school” OR “elementary school” OR “junior school”) AND (“active transport” OR “active commute” OR “active travel”)3071
Transport Research International Documentation1: schoolchild OR child
2: school OR primary school OR elementary school OR junior school
3: active transport OR active commute OR active travel OR walk OR bike OR cycle OR bicycle OR skateboard OR vehicle OR motorised transport OR car OR drive
4: 1 AND 2 AND 3
176
Google Scholar(schoolchild OR child) AND (“school” OR “primary school” OR “elementary school” OR “junior school”) AND (“active transport” OR “active commute” OR “active travel”) AND (walk OR bike OR cycle OR bicycle OR skateboard OR vehicle OR “motorised transport” OR car OR drive)18,600
Table 2. Inclusion and exclusion criteria used for the literature review.
Table 2. Inclusion and exclusion criteria used for the literature review.
Inclusion CriteriaExclusion Criteria
  • Publication year: Published before March 2023
  • Article type: Investigative studies (quantitative or qualitative), systematic reviews and synthesis, commentaries, editorials, and theses
  • Target age group: Between 6 and 12 years
  • Study focus: Address any of the 7 key issues pertaining to ASC below:
    • What are the trends of the global prevalence of ASC?
    • How does ASC contribute to the physical activity levels of school children?
    • What are the health benefits of ASC?
    • What are the determinants of ASC behaviours?
    • What interventions are effective in increasing ASC?
    • What are the advancements in research techniques related to ASC?
    • What are the current limitations in ASC research?
  • Non-English literature
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lam, H.Y.; Jayasinghe, S.; Ahuja, K.D.K.; Hills, A.P. Active School Commuting in School Children: A Narrative Review of Current Evidence and Future Research Implications. Int. J. Environ. Res. Public Health 2023, 20, 6929. https://doi.org/10.3390/ijerph20206929

AMA Style

Lam HY, Jayasinghe S, Ahuja KDK, Hills AP. Active School Commuting in School Children: A Narrative Review of Current Evidence and Future Research Implications. International Journal of Environmental Research and Public Health. 2023; 20(20):6929. https://doi.org/10.3390/ijerph20206929

Chicago/Turabian Style

Lam, Ho Yeung, Sisitha Jayasinghe, Kiran D. K. Ahuja, and Andrew P. Hills. 2023. "Active School Commuting in School Children: A Narrative Review of Current Evidence and Future Research Implications" International Journal of Environmental Research and Public Health 20, no. 20: 6929. https://doi.org/10.3390/ijerph20206929

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop