The recovery model in chronic mental health: A community-based investigation of social identity processes

Affiliations.

  • 1 Research School of Psychology, The Australian National University, Canberra, Australia. Electronic address: [email protected].
  • 2 School of Psychology, The University of Queensland, Brisbane, Australia.
  • 3 School of Public Health, The University of Queensland, Brisbane, Australia.
  • 4 ANU Medical School, The Australian National University, Canberra, Australia.
  • PMID: 32590231
  • DOI: 10.1016/j.psychres.2020.113241

The recovery model has been enormously influential in shaping mental health services globally over the last two decades. However, empirical research on its outcomes and psychological mechanisms is sparse. This community-based case study utilised both semi-structured qualitative interviews and quantitative survey methods to investigate perceptions of recovery, identity, and wellbeing among people with chronic and severe mental illness attending recovery-oriented support groups. Consistent with a social identity approach and the recovery model, to the extent that people identified as "in recovery", they reported better recovery outcomes (e.g., sense of purpose) and reduced psychological distress. Furthermore, recovery identity more strongly predicted recovery outcomes than it did psychological distress. Both the quantitative and qualitative data pointed to collective efficacy (i.e., group-based empowerment) as a key mediator of these outcomes. These findings are consistent with the recovery model and speak to the utility of a social identity approach for conceptualizing its efficacy. However, these findings also speak to the need for further evaluation of how and when recovery-oriented mental health services achieve their intended goal of improving quality of life for people with chronic and severe mental illness.

Keywords: Chronic mental illness; Collective efficacy; Mental health recovery; Social identity; Well-being.

Copyright © 2020 Elsevier B.V. All rights reserved.

Publication types

  • Research Support, Non-U.S. Gov't
  • Case-Control Studies
  • Chronic Disease
  • Community Mental Health Services / methods*
  • Mental Disorders / psychology*
  • Mental Disorders / therapy*
  • Mental Health / trends*
  • Mental Health Recovery*
  • Middle Aged
  • Quality of Life / psychology
  • Residence Characteristics
  • Social Identification*
  • Young Adult
  • Open access
  • Published: 17 January 2017

An Integrated Recovery-oriented Model (IRM) for mental health services: evolution and challenges

  • Barry G. Frost 1 , 2 ,
  • Srinivasan Tirupati 3 , 4 ,
  • Suzanne Johnston 3 ,
  • Megan Turrell 3 ,
  • Terry J. Lewin 2 , 3 , 4 ,
  • Ketrina A. Sly 2 , 3 , 4 &
  • Agatha M. Conrad 2 , 3 , 4  

BMC Psychiatry volume  17 , Article number:  22 ( 2017 ) Cite this article

36k Accesses

61 Citations

11 Altmetric

Metrics details

Over past decades, improvements in longer-term clinical and personal outcomes for individuals experiencing serious mental illness (SMI) have been moderate, although recovery has clearly been shown to be possible. Recovery experiences are inherently personal, and recovery can be complex and non-linear; however, there are a broad range of potential recovery contexts and contributors, both non-professional and professional. Ongoing refinement of recovery-oriented models for mental health (MH) services needs to be fostered.

This descriptive paper outlines a service-wide Integrated Recovery-oriented Model (IRM) for MH services, designed to enhance personally valued health, wellbeing and social inclusion outcomes by increasing access to evidenced-based psychosocial interventions (EBIs) within a service context that supports recovery as both a process and an outcome. Evolution of the IRM is characterised as a series of five broad challenges, which draw together: relevant recovery perspectives; overall service delivery frameworks; psychiatric and psychosocial rehabilitation approaches and literature; our own clinical and service delivery experience; and implementation, evaluation and review strategies. The model revolves around the person's changing recovery needs, focusing on underlying processes and the service frameworks to support and reinforce hope as a primary catalyst for symptomatic and functional recovery. Within the IRM, clinical rehabilitation (CR) practices, processes and partnerships facilitate access to psychosocial EBIs to promote hope, recovery, self-agency and social inclusion. Core IRM components are detailed ( remediation of functioning; collaborative restoration of skills and competencies; and active community reconnection ), together with associated phases, processes, evaluation strategies, and an illustrative IRM scenario. The achievement of these goals requires ongoing collaboration with community organisations.

Conclusions

Improved outcomes are achievable for people with a SMI. It is anticipated that the IRM will afford MH services an opportunity to validate hope, as a critical element for people with SMI in assuming responsibility and developing skills in self-agency and advocacy. Strengthening recovery-oriented practices and policies within MH services needs to occur in tandem with wide-ranging service evaluation strategies.

Peer Review reports

Disorders such as schizophrenia were historically viewed as chronic, degenerative illnesses, with little prospect of improvement or recovery. These negative and debilitating notions of serious mental illness (SMI) were challenged by the consumer movement, with recovery perspectives bringing a new sense of meaning and purpose to individual’s lives, even though symptoms might remain [ 1 – 4 ]. However, in the absence of clear operational or scientific definitions of ‘recovery’, it was questioned whether the process would be understood and amenable to collaborative interventions [ 5 ], or the value of the term compromised [ 6 ] and potentially commandeered by those seeking to reduce service costs [ 7 , 8 ]. Concerns that recovery-focused initiatives could default to rhetoric rather than practice were also raised [ 9 , 10 ].

Consumer research identified recovery as both a process and an outcome, involving factors related to personal wellbeing and social inclusion, which were distinct from traditional clinical domains [ 4 , 11 ]. Nevertheless, some scepticism remains around the notion of recovery [ 12 ], coupled with concerns that the burden of risk will be borne by families and carers [ 13 ]. It is generally accepted that improved mental health (MH) outcomes can be achieved through access to a range of psychosocial evidence-based interventions (EBIs) [ 10 , 14 – 16 ]. However, sufficient service ‘infrastructure’ needs to be activated to ensure recovery-oriented approaches are successfully embedded into everyday practice and access to EBIs is enhanced.

Advances in psychopharmacology made it possible for many people with SMI to be discharged from long-stay care. However, they were often discharged from highly structured inpatient environments with little provisioning for their needs, which according to some reports, did not extend beyond a prescription [ 17 ]. It became increasingly apparent that many individuals experience a constellation of signs and symptoms superimposed and interacting with a background level of impairment and disability. Function is often impaired across multiple domains (e.g., cognition, living skills, social skills, occupation/education) and the level of impairment can often be exacerbated by relapse and deteriorate further with subsequent episodes [ 15 ].

Whilst psychopharmacological treatments have improved and are considered fundamental to illness management, their role in the restoration of skills considered essential for a satisfying and fulfilling life is at best limited. For example, Meltzer [ 18 ] was unable to identify a correlation between amelioration of positive symptoms and social outcomes. It is also evident that medications have not solved the problem of relapse [ 19 ] and carry significant side effects and risks, including over-reliance, poly-pharmacy and inappropriate use [ 20 ]. Following a 20 year longitudinal study, Harrow et al. [ 21 ] state that “ antipsychotics are not effective in eliminating or reducing psychosis for the great majority … and may impede recovery of some … ” (p. 3013). Further, Deacon [ 22 ] suggests that under biomedical treatment models there has been a sharp increase in psychiatric medication use, a broad lack of clinical innovation, and poor MH outcomes.

Despite calls for reform, the disparity between the recovery needs of individuals with SMI and service delivery paradigms is reflected at several levels. For example, among young people, schizophrenia remains one of the top ten causes of disability [ 23 ]. People with psychotic disorders represent 25% of total disease burden [ 24 ] and schizophrenia is the 3rd most important disease in terms of years lived with disability for those aged 15–44 years [ 25 ].

Poor physical health is also experienced by many people with psychotic disorders, with 45.1% classified as obese and 33.5% assessed as having low physical activity [ 26 ]. The majority of people with SMI are also unemployed (78.5%), have poor education levels, impaired social skills (63%), and limited contacts [ 26 ]. Consequently, the estimated annual economic cost in Australia for all psychotic disorders is $4.91 billion from a societal perspective and $3.52 billion from a government perspective [ 27 ]. Moreover, even though annual costs have been relatively stable (over the 2000–2010 decade), there has been a significant redistribution of costs to the non-health sector, in line with Australian government initiatives [ 28 ]. The high and continuing levels of burden associated with SMI have prompted some authors to call for ‘widespread systemic change’ to MH systems, promoting an increased emphasis on shared decision making, independence (e.g., financial, residential, personal) and social connectedness [ 10 ].

Regional opportunities and imperatives

Like many countries, Australian MH services are currently in a state of transition, including: formulation of national frameworks with an increased focus on recover-oriented care provision [ 29 – 31 ]; development of a new Australian MH Care Classification [ 32 ]; and introduction of Activity Based Funding [ 33 ]. In broad terms, recovery-oriented service delivery : “ … is centred on and adapts to people’s aspirations and needs, rather than people having to adapt to the requirements and priorities of services ” and it has a “… responsibility to provide evidence-informed treatment, therapy, rehabilitation and psychosocial support that assist in achieving the best outcomes for people’s mental health, physical health and wellbeing ” [ 30 ], p. 26.

Within New South Wales (NSW), planning commenced in 2005 to establish a number of sub-acute inpatient MH units, with the primary goal of improving access for people with SMI to recovery-focused rehabilitation services that were highly integrated and rigorously evaluated [ 34 ]. This provided an opportunity for Hunter New England Mental Health services to develop an innovative model of care at a level of service delivery that had not previously been explored. Details about the specific 20-bed, sub-acute Intermediate Stay Mental Health Unit (ISMHU) that was initially established are provided elsewhere, together with our preliminary service evaluation [ 35 ].

Importantly, development of a new level of regional MH care necessitated consideration of all of the potential MH service pathways and partnerships, together with their treatment models and intended goals. Within this context, and given the limited availability of established service-wide, recovery-focused models of care [ 36 – 39 ], a broader framework for an Integrated Recovery-oriented Model (IRM) for MH services was formulated, which sought to support and promote ‘ remediation , restoration and reconnection ’.

The primary purpose of this paper is to outline the IRM and to stimulate ongoing refinement of recovery-oriented service models. Evolution of the IRM is characterised with respect to five broad challenges. The first three challenges relate to identification of: 1) relevant recovery perspectives; 2) overall service delivery frameworks and models; and 3) key features and processes associated with current specialised clinical rehabilitation (CR) interventions for people with enduring SMI. The fourth, or central challenge, is to draw together the main elements from these first three challenges into a coherent, service-wide IRM for MH service delivery. The fifth challenge relates to devising relevant implementation, evaluation and review strategies for recovery-oriented MH service models and components.

Recovery perspectives

Challenge 1.

Identifying the aspects of personal and clinical ‘recovery’ and related approaches that need to be considered in re-designing ‘recovery-oriented’ MH services.

Recovery possibilities and needs

Research has shown that recovery is possible [ 40 , 41 ] and that people with SMI value the opportunity to participate and contribute to society [ 42 ]. However, for many there is limited access to EBIs that may prove effective in supporting hope and restoring confidence and competence [ 43 ]. Mojtabai et al. [ 44 ] found that more than 50% of people with schizophrenia received either no treatment or suboptimal treatment. Torres-González et al. [ 45 ] identified six areas of specific need: frequent complications and co-morbidities (e.g., substance misuse); psychological, social and economic needs; early interventions to reduce illness progression; treatment augmentation with rehabilitation EBIs; maintenance of service contacts; and greater research efforts into existential needs. Better access to psychosocial interventions and well-managed medication are warranted [ 14 , 45 ], together with a shift away from case/risk management practices to service models that facilitate access to EBIs [ 10 , 20 ].

Recovery goals

The term recovery is clearly multi-layered. Nevertheless, it carries an unequivocal message of a better outcome, conveying a sense of hope; it may also carry expectations in regard to interventions, timeframes and supports. Attempts to reintroduce hope and optimism are based on the view that recovery is possible even though residual limitations may remain. Unlike physical medicine, where recovery goals are generally well understood, the role and significance of rehabilitation for people with SMI has been less well understood - even though psychiatric rehabilitation has always been about ‘recovery’ [ 19 ] and supporting self-determination and independence through improvements in wellbeing and role functioning.

Snyder et al. [ 46 ] described hope as “ the person’s perceived ability or internalised belief that he or she can produce goals, pathways and agency ” (p. 89), suggesting that, as a goal directed motivational process, hope requires constant feedback and agency. If hope is a catalyst for change and improved health outcomes [ 46 ], the question arises as to how hope is both generated and sustained. This also brings into focus the ethical requirements of beneficence (doing good) and maleficence (avoiding harm) that typically guide health service provision. Some recovery-oriented frameworks propose that hope may be generated through service and cultural reforms; for example, “… the physical, social and cultural service environment inspires hope, optimism and humanistic practices for all who participate in service provision ” ([ 47 ], p. 7). Although such statements are very positive, they run the risk of being overpowered and reverting to rhetoric, unless driven by outcomes that reconfirm the considerable investments in recovery.

Le Boutillier et al. [ 37 ] suggested that promoting citizenship and a clear sense of place are core goals for recovery-oriented MH services, the primary purpose of which is to encourage self-agency. Validating personal goals can also help to reduce a client’s sense of frailty and hopelessness. Liberman and Kopelowicz [ 5 ] proposed that as improvements are made in a range of personally valued domains, more subjective qualities such as hope, empowerment and autonomy become evident. Snyder et al. [ 46 ] suggested that the processes of hope and rehabilitation “ fuel each other in an iterative manner over the temporal course of treatments ” (p. 107). Recovery can be complex and non-linear, with hope seen as critical in shaping and sustaining improvements in a range of skill domains, consistent with social inclusion [ 48 , 49 ].

Recovery processes

Early access to rehabilitation interventions has been associated with better functional outcomes [ 50 ]. Making rehabilitation available across the continuum of care may reduce health costs by shortening hospital admissions, reducing activity limitations, and improving quality of life. More generally, the discipline of psychiatric rehabilitation has contributed much to improving service delivery and outcomes [ 17 ]. Psychiatric rehabilitation challenged the MH system to think more expansively and respectfully about people with SMI, promoting choice, shared decision-making, consumer involvement, and a focus on inherent strengths and recovery possibilities.

The discipline of psychiatric rehabilitation promoted the adoption of a broad, holistic approach and advocated for access to quality residential, education and employment opportunities. Quality frameworks were also introduced, including comprehensive multidisciplinary and inter-service team reviews. Due to the obvious synergies with the recovery approach, rehabilitation services have been proactive in adopting consumer oriented recovery strategies. Much has also been done to reduce the negative approach associated with the official nosology of schizophrenia, in which therapeutic nihilism and stigma have operated as self-fulfilling prophecies [ 5 ]. Perhaps, reluctance to accept the discipline stems from the fact that psychiatric rehabilitation is relatively easy to define but, as highlighted by Anthony and Farkas [ 17 ], any explanation belies the complexities of the processes involved.

An understanding of personal recovery as a subjective experience has emerged and this meaning now underpins MH policy internationally e.g., [ 38 , 51 ]. While the provision of recovery-oriented care is a guiding principle, implementing recovery-oriented or recovery-enabling [ 52 ] practices requires transformations within MH systems [ 10 , 38 , 39 ]. In some sectors, such as MH inpatient settings, there is limited research directly addressing recovery-oriented practice [ 39 , 53 ]. However, a recovery enabling framework has been proposed to address workforce gaps in core recovery competencies among inpatient providers [ 52 ].

Until recently, the focus was almost exclusively on clinical recovery [ 54 ]. Central to the delivery of recovery-oriented services is a shared understanding of recovery between consumers, carers and health professionals [ 51 ]. Recovery-oriented psychiatric rehabilitation can be seen as supporting people with SMI in the pursuit of a meaningful life [ 55 ]. As recovery is an ongoing and non-linear process, recovery-oriented experiences and opportunities during periods of hospitalisation also need to be adequately addressed [ 52 ].

Recovery contexts

Once again, it needs to be explicitly acknowledged that recovery experiences, opportunities, trajectories, and evaluations are inherently personal. Among people with SMI, recovery is generally viewed as “ a journey of small steps ”, within which participation in everyday activities is “ frequently considered as both facilitators and indicators of recovery ” ([ 10 ], p. 237). Moreover, while the current paper is primarily about recovery-oriented MH service provision, there are a broad range of potential recovery contexts and contributors and, for many people, professional interventions may play a relatively minor or time-limited role [ 56 ]. On the other hand, individuals with enduring SMI are likely to be influenced proportionately more by the attitudes and practices of specialised MH, general health, and community managed services. Importantly, key processes associated with recovery (e.g., sustaining hope, promoting self-agency and reconnection) need to occur both within and outside of MH services [ 56 ] and, where possible, be enhanced by integrated, recovery-oriented practices.

Service delivery frameworks and models

Challenge 2.

Reconciling the broad array of general and specialised service delivery frameworks, models and intervention strategies of potential relevance to ‘recovery-oriented’ MH services.

There are numerous recommendations about service delivery approaches, ranging from general health or MH focused over-arching ‘frameworks’, through broad ‘intervention strategies’ or ‘models’, to specific ‘targeted interventions’. The WHO International Classification of Functioning, Disability and Health [ICF, 57] provides a general framework for considering the spectrum of needs of people with SMI. Integrating medical and social models for people with health conditions, the ICF focuses on human functioning, activity and participation, rather than disease and disability. It also provides a comprehensive guide to the identification of a range of protective and risk factors. For example, at the level of body function, the ICF framework includes consideration of psychotic symptoms, poor concentration and memory, low self-esteem and confidence. Activity limitations may include poor self-care, poor physical health, social withdrawal, and an inability to follow instructions. Participation restrictions may be reflected as the inability to continue education, difficulties maintaining social relations, problems with accommodation and accessing recreational activities. Consequently, an array of recovery-oriented approaches may be required to promote and sustain hope and resilience, facilitating improvements in personal functioning, activity and social participation.

The ICF has previously been implemented in an Italian psychiatric rehabilitation setting and reported to be a helpful framework among people with SMI, promoting a common language and integrated treatment model, supporting the development of client focused individual rehabilitation plans and improving services [ 57 ]. Similarly, individualised approaches to recovery in vocational rehabilitation have found positive effects on both clinical and employment outcomes [ 58 ]. Although the research literature provides some assistance in regard to recovery-oriented frameworks, it provides limited guidance on optimal delivery systems or recovery-oriented models for MH services.

Perkins and Slade ([ 59 ], p. 33) noted that “ there can be no ‘blueprint’ for recovery – each person must find their own way ”, although key factors important in supporting recovery-oriented practice and transforming MH services have been identified in the recovery literature. Le Boutillier et al. [ 37 ] proposed a conceptual framework to guide practice, focusing on four domains: promoting citizenship; organisational commitment; supporting personally defined recovery; and working relationships. Hopper [ 60 ] viewed recovery as a therapeutic endeavour and proposed four stages in the recovery process: renewing a sense of possibility; regaining competencies; reconnecting and finding a place in society; and reconciliation. Rodgers et al. [ 61 ] employed a staged approach, mapping EBIs for each stage of the recovery process.

From a service model perspective, Thornicroft and Tansella [ 62 ] suggested service configurations should be balanced between hospital and community services, outlining three levels of care: primary care with specialist back-up; mainstream MH care; and specialised MH services. Specialised services included: early intervention; assertive treatment teams; alternatives to acute inpatient care; residential care and vocational rehabilitation. Adopting a slightly different approach, Flannery et al. [ 63 ] developed a service model based on the core functions required for a recovery-focused MH system: acute care (community teams and alternatives to inpatient care); emergency services; continuing care partnerships (assertive treatment teams, supported accommodation, therapy services, vocational rehabilitation and drop-in centres); and early intervention services. Although this pragmatic approach could be introduced with minimal cost, it is unclear how access to EBIs and other major requirements of recovery-focused models would be achieved. The fundamental tenant of any reform should be that recovery is supported as both a process and an outcome. If this does not occur, there is an inherent risk that traditional imperatives will prevail and re-establish a disconnected dichotomous system (e.g., acute/emergency vs. disability support services).

Slade et al. [ 38 ] identified ten validated interventions that support recovery by targeting key processes (connectedness, hope, identity, meaning and empowerment [CHIME]) [ 64 ], illustrative of the types of interventions expected in recovery-oriented MH systems. These included: peer support workers; advance directives (if future capacity is lost); wellness recovery action planning (WRAP) tools and processes [ 65 ]; illness management and recovery (IMR) [ 66 ]; the REFOCUS model (recovery-promoting relationships and work practices) [ 67 , 68 ]; strengths-based models [ 69 ]; recovery colleges or recovery education programs; individual placement and support (IPS) [ 70 ]; supported housing; and MH trialogues (community forums). Many of these EBIs can be implemented regardless of the specific recovery-oriented model; although some have been evaluated predominantly in community MH settings [ 38 ]. Others involve more complex manualised pro-recovery interventions or modules, such as the REFOCUS model, IMR program, and WRAP, which also emphasises peer support in the development of individual recovery plans [ 65 ]. Strengths-based case management models supporting consumer directed care have also been implemented in both acute and community MH settings [ 69 , 71 ], focusing on personal strengths and goals rather than deficits, and integrating a variety of EBIs. While all approaches support recovery, few provide an overarching framework and service-wide model for MH care provision.

Internationally, implementing recovery-oriented practices has posed challenges for MH services [ 65 , 72 ]. In Australia, a need for MH systems transformation has also been identified, in order to provide a continuous recovery-oriented care framework that links acute inpatient and community services [ 73 ]. Recent conceptualisations of recovery-oriented practice have focused primarily on clinical and personal recovery; however, a new concept of service-defined recovery is seen as translating recovery into practice according to the goals and needs of an organisation [ 74 ]. This accords with earlier suggestions that an ideal model should “ link the abstract concepts that define recovery with specific strategies, that systems, agencies and individuals can use to facilitate it ” ([ 75 ], p. 482). While service approaches operationalising recovery-oriented practice are yet to be extensively evaluated, research on staff perspectives has identified perceived barriers (e.g., competing priorities in providing recovery-oriented support), which also highlight the need for a whole-systems approach in transforming services [ 74 , 76 ].

Clinical Rehabilitation (CR) within MH services

Challenge 3.

Building on the core elements of psychosocial and MH rehabilitation, to facilitate service provision along a recovery-oriented continuum, with specialised clinical rehabilitation processes and services nearer to one extremity, delivering targeted MH interventions and supporting people with enduring SMI.

In part, we use the expression ‘CR within MH services’ to draw a distinction with ‘disability support’ (associated primarily with care linked to enduring functional impairment or other activity limitations ) and to de-emphasise the discipline-specific aspects of ‘psychiatric rehabilitation’, in favour of a recovery-oriented care continuum of relevance to all MH workers. All of these approaches have roles to play but require different skills sets, competencies and professional and clinical processes. Encouraging clients to progressively assume independence and responsibility for their own care is axiomatic to CR and consistent with personal recovery approaches [ 6 , 16 , 17 , 19 , 77 ]. Given that CR provides a unique opportunity to empower people with SMI to assume greater levels of self-agency, the question arises as to how these opportunities can be further realised within service delivery models that not only respect this role but also complement and enhance opportunities for recovery and social inclusion?

CR employs a set of interventions and processes that aim to achieve and maintain optimal functioning in the client’s environment of choice. CR is about helping individuals to realise their personal goals, in a supportive context that builds trust and confidence in self-agency. It is about affirming and reaffirming that the investment of hope in personal coping and everyday functional skills has been justified and, in so doing, support the independent exploration of new and more satisfying personal goals.

Developing interventions and supports that promote recovery and challenge commonly held stereotypes, which by definition disable and segregate, is a complex undertaking. Hope is a key factor in this process and, in taking the first tentative steps to regaining a sense of control and self-agency, it is vitally important to understand the risks involved and to ensure trust and personal dignity are protected. Ensuring that an individual’s investment in rehabilitation and recovery processes is supported, and not adversely affected as new goals are explored, is also critical.

Depending on individual recovery goals, CR may involve single or multiple EBIs delivered by a skilled practitioner, in conjunction with a CR team. The interventions should be developed in a collaborative, empowering and optimistic manner, based on a thorough understanding of the person’s goals and abilities (including both strengths and vulnerabilities). The plan may also be cross-sectoral, involving health professionals working in conjunction with general practitioners (GPs), community support agencies, as well as educational, employment and housing organisations. From a service-led recovery perspective [ 74 ], it should also be recognised that there may need to be different service streams even within specialised CR services, reflective of variations in the complexity of client needs and available resources; for example, some service streams may offer targeted, time limited EBIs, while others provide more of a ‘continued care’ approach, supporting clients with enduring SMI to maintain their MH and community tenure.

CR principles and priorities

Foremost among the key features of CR are the principles that guide the delivery of recovery-focused interventions: recovery-oriented; promoting independence; person-centred; flexible, responsive and inclusive; accommodating different learning styles; focusing on strengths; utilising EBIs; providing integrated multidisciplinary care (including service continuity); and facilitating community and environmental supports. Some of the CR processes that flow from these principles are detailed in Table  1 , including establishing recovery-oriented goals, undertaking assessments and recovery planning, delivering interventions, and clinical review or recovery-focused tracking.

People change and grow, and various factors promote positive adaptation, such as setting your own goals, learning new skills, hope, and self-efficacy [ 17 ]. With respect to specific or targeted CR intervention priorities, Mueser et al. [ 16 ] recently classified psychosocial interventions according to whether the evidence was sufficient or promising. Included among the established EBIs were: cognitive behavioural therapy for psychosis; cognitive remediation; family psycho-education; illness self-management training; social skills training; and supported employment. Other interventions considered to be very promising [ 16 ] were: social cognitive remediation [ 78 ]; cognitive adaptive training [ 79 ]; integrated psychological therapy [ 80 ]; healthy lifestyle interventions [ 81 ]; and supported education [ 82 ]. Additional interventions with an evidence base included: motivational interviewing reviewed by [ 83 , 84 ]; errorless learning [ 85 ]; skill building reviewed by [ 86 ]; and family interventions reviewed by [ 87 ].

Specialised CR services may also require a staffing compliment and roster arrangements that depart from traditional approaches. Ideally, staff should be recruited against a set of values and competencies consistent with rehabilitation and recovery-oriented approaches, including: openness; empathy and encouragement; supporting responsible risk taking; a positive outlook; a collaborative focus on client’s inner resources and strengths, and a preparedness to go the extra distance [ 88 ]. Experience suggests that CR staff also need to be patient, resourceful, and innovative, and enjoy problem solving. Professional background and training is also important, as some professions have extensive theoretical and practical training in provision of complex interventions. For example, increasing the number of occupational therapists, social workers and psychologists, relative to those with generalist training, may significantly increase service capacity and recovery focus. However, such guidelines may be misleading, as some generalist-trained staff with a passion for CR may make outstanding contributions. Importantly, CR teams should also include consumer advocates, as these staff may provide direct assistance to clients and clinical staff, and help ensure that the team retains a strong client-centred recovery-oriented approach.

Integrated Recovery-oriented Model (IRM)

Challenge 4.

Developing a recovery-oriented model for MH service delivery (promoting ‘remediation, restoration and reconnection’) that provides both an overarching, inherently collaborative and integrated approach, together with identification of opportunities for targeted specialist CR initiatives.

The IRM was designed to support the recovery needs of people with SMI by improving access to a range of EBIs provided within a service context that reinstates hope, rebuilds competencies and provides opportunities to reconnect. Three foundation elements or functions of this service model that partner with the individual client include: acute/emergency MH care; specialised CR; and community managed/non-government organisations (CMOs/NGOs) providing community integration services .

The IRM operates as a tripartite agreement, with each of the partners providing recovery-focused services in an integrated and seamless manner. Each of the core services may also operate in conjunction with a range of other specialist services (e.g., sub-acute inpatient, substance misuse, neuropsychiatry) and community-based organisations, including GPs, accommodation services, employment services, education providers, drop-in centres, community participation and recreation services. To ensure continuity, the IRM requires flexibility, transparency and responsiveness, but with the degree of service involvement titrated according to client recovery needs. Clearly, this requires a solid understanding by all partners of service and management core functions and processes. Consequently, a major strength of the IRM is the ability to safeguard hope and self-esteem by intervening early to preserve coping and functional skills across a number of domains, including everyday living skills, accommodation, social networks, employment and education endeavours.

Key principles guiding service delivery within the IRM include: 1) services are recovery-oriented; 2) care delivered is person-centred, holistic and inclusive; 3) care enables and supports choice and self-management; 4) services are integrated across the care continuum; 5) service delivery is seamless and complementary across all providers (i.e., no ‘wrong door’); 6) services and care are based on the most appropriate available evidence; 7) partnerships with other services, government departments and CMOs/NGOs are integral to service delivery; 8) consideration of equity issues informs decisions about services and care; 9) information technologies are used to improve access to care, facilitate enhanced collaboration and communication within the service, consumers, their families and carers; and 10) services and care delivery is aligned with national, state and local directions.

The three main components of the IRM have been based on the ICF concepts of function, activity and participation [ 89 ], but also incorporate elements identified by Hopper [ 60 ]. Under the IRM, it is proposed that acute services should focus on ameliorating positive symptoms and reinstating a sense of possibility. At the earliest available opportunity, CR services, supported by CMOs/NGOs, would begin to restore hope through the development of a range of skills pertinent to personal goals. As the client regains confidence, CMO/NGO services would focus on exploring opportunities that would reinforce personal recovery and reconnection with the community. However, it also needs to be acknowledged that there is variation across Australian States in the service delivery roles performed by CMO/NGO services, and even more so from an international perspective. The manner in which these remediation, restoration and reconnection components revolve around the person's changing recovery needs is highlighted in Fig.  1 . The overlapping and, somewhat idealised, sequential phases of recovery are further illustrated in Fig.  2 ; acknowledging again that recovery can be multi-layered and non-linear [ 48 , 49 ]. More detail about the complementary roles of the respective IRM components is provided below.

Integrated Recovery-oriented Model (IRM) for mental health services

Integrated Recovery-oriented Model (IRM) - Phases of recovery

Remediation of functioning - reinstating a sense of possibility

This phase is the start of a complex journey in which the key elements that generate and sustain hope must be carefully reintroduced and nurtured. The goals are to intervene early to reduce the psychological and social sequelae associated with the onset of illness. Building trust and hope that is real and sustainable will be critical in developing a positive adjustment to the diagnosis. This phase also provides an opportunity to address physical health issues, ensure safety, manage any legal and financial issues, and to identify other likely impacts on the person, their partners, families and friends. When a person’s coping and protective strategies have been breached, resulting in acute psychosis, they are likely to feel overwhelmed, shocked, confused, fearful, anxious, in denial and exhausted. These reactions may be fuelled by stigma and run the risk of being exacerbated by treatment and management plans that are: circumspect in their vision; fail to respect and value the person, their family’s needs and aspirations; or lack credibility in terms of delivery and coordination.

Initial treatment provides an invaluable opportunity to reduce fear associated with the onset of symptoms and the diagnosis, and to commence development of a collaborative recovery-oriented plan that is consonant with the wishes and aspirations of the person and their family. To ensure that the client’s investment of hope is well placed, it is essential that there is a full understanding of their strengths, protective factors and possible risks. As with physical rehabilitation, care needs to be exercised as the events and triggers that precipitated the relapse are brought into sharp focus by an approaching discharge. The need for care is also reinforced by the knowledge that a successful resolution of positive symptoms does not necessarily indicate a return to pre-episode functioning. A thorough assessment is required to develop a supportive, individually tailored, multi-modal skill building program, which may be provided in combination with other treatments; a point highlighted in a recent review by Lyman et al. [ 86 ].

The need for a holistic plan, which supports hope through a range of strategies that build confidence and competencies and addresses vulnerabilities, underscores the importance of the early involvement of rehabilitation specialists. While this phase will generally be led by acute MH services (which have specific expertise in treating positive symptoms), they also require the support of CR, and CMO/NGO services, in building confidence and hope in a plan that extends beyond the acute setting. In order to demonstrate an unequivocal commitment to the goals of the collaborative recovery-oriented plan, a number of relevant clinical and nonclinical services may need to be involved, including: emergency assessment and triage; acute inpatient and community services; community MH teams; early intervention programs; specialist clinicians; and associated links with GPs, sub-acute inpatient and other specialist agencies.

Restoration – enabling, regaining competencies

The goal of this phase is to demonstrate that hope and the sense of possibility are valid constructs in the pathway to recovery. At the earliest opportunity, a range of EBIs should be available to assist in rebuilding or confirming personal, interpersonal and daily coping skills and competencies. This may also provide an opportunity to redress developmental gaps and lifetime goals, both of which could contribute to a renewed sense of self. As confidence is developed in personal coping skills and environmental adaptations, a more robust foundation for further pathway or goal-directed thinking should emerge. Exploring new and confirmatory experiences will obviously entail a degree of positive risk taking and comprehensive strategies may need to be in place to safeguard personal dignity. Throughout this phase, the focus will be unequivocally on the development of self-agency, particularly as it relates to mental and physical recovery, and social inclusion.

For some people with SMI, the recovery journey may initially hold few protective factors and pose considerable challenges and risks. For example, a move from a highly structured inpatient unit to a loosely structured home or residential setting, with a questionable and fragile confidence in coping skills and supports, may pose major risks. Insufficient supports during this challenging period may propel a person to find membership in segregated company or attempting to self-manage through the use of non-prescribed substances. Transitional arrangements may provide an opportunity to build confidence and minimise stress, as well as providing a positive foundation on which to build essential psychological and everyday functional skills. The development of additional competencies may include: strategies to manage residual symptoms; cognitive skills; social skills; activities of daily living; physical health; family education and support; and supported education or employment. These interventions should be based on a comprehensive assessment, including usage of collaborative tools such as the Mental Health Recovery Star MHRS; [ 90 ], and a collaboratively developed recovery-oriented plan.

CR services need to work in partnership with acute services, both inpatient and community, and CMOs/NGOs, but without duplicating either. CR services should be most closely aligned with community-based services, both clinical and non-clinical. Given the multitude of factors impacting on recovery, there is no single formula with which to predict or determine outcomes and timeframes [ 59 ]. For example, within non-acute MH services the timeframe for full client engagement would typically be for a period up to 12 months, but the overall extent of involvement, including partial or backup clinical support, would be dependent on a range of individual, social and environmental circumstances. This phase should be led by CR services but with significant involvement of CMOs/NGOs and back-up from acute and emergency services. The potential service elements include: CR teams and streams supporting both targeted and continuing care roles; specialist CR interventions; intermediate (sub-acute) stay recovery units – step-up and step-down; and links with early intervention services, GPs, housing providers, employment, education and other non-acute inpatient services.

Reconnection - with place and society

The aim of this phase is to reconnect and re-establish a place in the community, and to explore opportunities for independence and social inclusion with a new sense of confidence and hope, based on the competencies developed in the previous stages. Development of a supportive daily structure is highly desirable, together with progressive utilisation and refinement of skills in the pursuit of a range of personal goals. This may necessitate graduated exposure to less structured or supported situations (e.g., independent living, community, social situations). During this phase, initial steps may be guided by CR clinicians but with CMO/NGO workers assuming greater responsibility as confidence grows in the client’s ability to be more independent. Essentially, this phase is about validating the investment of hope and developing greater levels of self-esteem and self-agency through exploration of opportunity.

One of the advantages of CMOs/NGOs lies in their capacity to build rich and full connections with other community based groups and services. These connections may open up many satisfying and life enriching opportunities for people with enduring SMI. CMOs/NGOs may assist in the exploration of these opportunities and in the development of: stable accommodation; civic and social activities, reducing social isolation; employment opportunities; recreational and sporting activities; as well as guidance in regard to relationships and existential needs. Importantly, many CMOs/NGOs have partnerships with GPs, which, together with initial support from CR services, may ensure better access and improved mental and physical health. This phase should be led by CMO/NGO services, with the level of input from CR titrated against personal recovery needs, clinical support, risk and legal issues. As the client becomes more confident in their self-determination abilities in the community, CR services should progressively withdraw, allowing the CMOs/NGOs to assume leadership. Acute MH services would always remain available for the transfer of care and joint clinical reviews. The potential service elements include: supported accommodation (low to very high residential); low support accommodation; day centres; links with GPs; specialist employment and education services; recreational and fitness centres; and home care services.

Illustrative IRM scenario

Application of CR planning, intervention, review, transfer and evaluation processes (detailed in Table  1 ) within the IRM, to support and promote recovery for an individual client, are illustrated in Fig.  3 . Examples of how the IRM can promote recovery for individuals with a SMI are, in most instances, complex but an illustrative scenario is provided in Table  2 . Here the remediation phase is characterised in terms of relapse prevention and admission related decisions designed to reinstate hope, while the restoration/reconnection phases are illustrated via a series of recovery-focused actions in response to different concerns (e.g., about medication, treatment/intervention adherence, coping strategies to manage stress, substance misuse, family dynamics, and safety).

Clinical Rehabilitation (CR) processes within the IRM supporting and promoting recovery

As an outline, this description of the IRM does not detail operational issues, such as: admission, referral and transfer processes; service hours; staff roles, competencies and training; service linkages; discharge pathways; and key performance indicators. Although these operational requirements should be guided by recovery-oriented and CR principles, other local and national factors may have an impact, including recording and reporting expectations. As with any reform, care also needs to be exercised in regard to agendas driven by vested interests and unrealistic expectations; most of all, there is a need to address the inertia within health services and to actively promote education and understanding of recovery and CR.

Evaluation and review

Challenge 5.

Devising implementation and evaluation strategies that enhance outcomes and facilitate review of recovery-oriented MH service models and components.

Evaluation goals – targets and perspectives

Evaluating the formulation, implementation and impact of specific intervention programs or MH service/practice changes can be a daunting task, especially when viewed from multiple stakeholder perspectives [ 91 ]. Expressed simply, the relevant issues are: what aspects of the service model are under evaluation (e.g., perceptions of practices and processes; EBI information, availability, uptake, fidelity and completion; compliance with guidelines and documentation; impact on clinical and/or personal outcomes; training and resource utilisation; and so on); from whose perspective (e.g., clients, carers, clinicians and/or service providers); with regard to what timeframes (e.g., initial impact, medium-term, ongoing); and using what evaluation methods or strategies (e.g., quantitative/qualitative, self-report, independent assessments, service data or other linkages).

The overriding question is: Can the chosen methods realistically address the identified evaluation goals within the required timeframes? In all likelihood, an assortment of evaluation strategies will be required, which vary in intensity and duration. Operationalising aspects of an evaluation could begin with a review of core resource materials and identified service pathways. For example, the 12 ‘clinical review’ items listed in the top right-hand corner of Table  1 could form the basis for a self-evaluation of CR processes within a particular service stream. Similarly, the flow diagram in Fig.  3 , which depicts IRM processes and phases, could provide a useful starting point for auditing progress for a sample of clients and identifying service barriers (e.g., evidence in clinical records of collaborative assessments and care planning, provision of EBIs, multi-disciplinary and interagency reviews).

Evaluation strategies – methods and measures

Ideally, program and service evaluations should incorporate a mixture of qualitative and quantitative methods, including: reviews of available evidence; client/carer/staff structured interviews or surveys; service audits; focus groups; first-person narratives and other feedback; and assessments of recovery trajectories, short- and longer-term outcomes, and associated processes and predictors. The latter could include: clinical recovery-focused evaluations (e.g., symptoms, medication compliance, relapse); personal recovery-focused evaluations (e.g., functioning, subjective wellbeing, independence and safety, social engagement, vocational activities, quality of life, community linkages); and service-related outcomes and evaluations (e.g., hospital presentations, contacts with community services, engagement/referral patterns, service transitions, staff perceptions and training, policy and guideline awareness, and associated costs).

There is a growing literature on the selection of strategies and measures for assessing recovery [ 4 , 92 – 94 ], recovery-oriented practice [ 65 , 74 , 95 ] and the recovery-orientation of services [ 71 , 92 , 96 , 97 ]. In choosing a particular set of tools, it may be useful to cover a representative range of recovery domains or processes, such as the CHIME spectrum described earlier [ 38 , 64 ] or the ‘broad superordinate recovery dimensions’ suggested by Whitley and Drake [ 98 ] (i.e., clinical, existential, functional, physical, and social dimensions of recovery). More generally, the capacity for client/carer self-evaluation of progress, for continuous review of recovery-oriented practices, and for reporting on key service outcomes and processes, need to become routine aspects of MH service provision, as recommended in several guidelines e.g., [ 31 ].

Preliminary local evaluations

While the IRM was developed as a service-wide model, it also provides an overarching framework for progressive MH service changes and EBI refinement. Like other programs [ 66 , 68 , 70 ], a staggered IRM introduction is probably more practical and likely to be endorsed; consequently, flexible, staged evaluation programs are also required. In our case, as described below, preliminary IRM-related evaluations focused on clinician perspectives (across the whole service) and IRM implementation within a purpose-built 20-bed, sub-acute unit [ 35 ]. More extensive evaluations are planned, covering a broader array of stakeholders and timeframes.

One staff-based method for evaluating variations in recovery-orientation is to survey clinicians pre- and post-service changes. For example, we surveyed MH clinicians recently, with the intention of conducting repeat surveys after full implementation of model of care changes. Preliminary findings ( N  = 251 clinicians, see Additional file 1 ) suggest that acute and community MH clinicians differ in their perceptions of the relevance of a range of recovery domains (e.g., social networks and work are perceived as less relevant domains by acute care clinicians), reflective of their likely differential contributions to the remediation and restoration phases of client recovery. Other studies have identified less positive attitudes towards recovery among inpatient providers [ 52 , 96 ], suggesting that treatment setting is an important factor to consider when refining recovery-oriented care practices and training.

Intrinsic to IRM evaluation and review is the ability to respond to new opportunities as they emerge and to ongoing feedback from various stakeholders. Our initial evaluation of the implementation of an IRM within a sub-acute ISMHU [ 35 ] provided preliminary confirmation that our 6-week recovery-oriented program was acceptable, valued, and capable of contributing to enhanced functioning and an improved recovery trajectory. With respect to the interface between program goals and the sensitivity of evaluation methods, within ISMHU the various EBIs (and associated program guides) were built around and expressed in comparable terms to the MHRS domains [ 90 ], the main collaborative assessment tool used within the unit, with marked admission to discharge MHRS improvements detected [ 35 ].

MH services have been the subject of many reforms but have remained largely disease-focused and paternalistic. The consumer lead recovery movement advocated for the adoption of more optimistic, recovery-oriented approaches, based on their experience that recovery was possible, despite residual symptoms [ 1 – 4 ]. Others have suggested that better access to treatments and psychosocial EBIs [ 10 , 14 – 16 , 43 ] is essential to improve overall MH outcomes, especially given the complexity of service and organisational reform.

Notwithstanding the merits of previous approaches, the reality is that EBIs are currently under-utilised and typically not delivered within sustainable, integrated MH systems. To assist people with SMI achieve their goals, service-wide frameworks for recovery-oriented care provision are clearly needed [ 36 – 39 ], together with validated intervention strategies and programs [ 16 , 38 ], and workforce education programs promoting recovery-enabling competencies and positive attitudes [ 52 , 74 ].

In this paper, we have drawn on relevant recovery perspectives, the psychosocial rehabilitation literature, and our own clinical and service delivery experience, to document the evolution of a broad IRM for MH services, together with associated challenges. A range of national [ 29 , 30 ] and State-led initiatives [ 34 ] to improve outcomes for people with SMI, including the establishment of intermediate stay units [ 35 ], provided a unique opportunity to explore recovery-oriented models of care. Based on ICF concepts [ 57 , 89 ] and CR principles [ 19 , 77 ], the IRM has attempted to address the broader recovery needs of people with SMI from a health rather than a disease perspective, and to view outcomes as the interaction between the health issues, the person and their environment.

It is easy to pigeonhole new service initiatives as simply ‘good clinical practice’. The IRM was developed to facilitate access to a range of recovery-oriented EBIs, integrated across the spectrum of need. The model focuses on the fundamental factors that have been shown to promote hope, recovery, self-agency, and social inclusion. The IRM includes evidence-based CR practices and processes as a substantive component (see Table  1 and Fig.  3 ), which have considerable potential to help realise individual goals and aspirations. However, as recovery is often complex and non-linear (see Fig.  2 ), the achievement of such goals is difficult in isolation, and requires the specialised contributions of acute, non-acute and community managed/non-government organisations (CMOs/NGOs).

Service delivery models such as the IRM encourage MH services to embrace opportunities to validate hope, as a critical element for people with SMI in assuming responsibility and developing skills in self-agency and advocacy. To promote ongoing refinement of recovery-oriented service models and inform policy development, wide-ranging evaluation strategies are also critical, some aspects of which have been briefly touched on in this paper, including some preliminary IRM related evaluations.

Importantly, the three core components of the IRM revolve around and interact with the person's changing recovery needs (see Fig.  1 ): remediation of functioning (directed towards reinstating hope and a sense of possibility); collaborative restoration of skills and competencies; and active community reconnection . These core components have equally significant roles to play in promoting recovery as a process and an outcome. The ‘ remediation, restoration, and reconnection ’ refrain also provides a convenient mnemonic for the broad types of support that should be expected from recover-oriented MH services. However, the inherent strength of the IRM lies not in the capabilities of each of the contributing specialties but in the potential of the tripartite collaboration to promote and sustain hope of a life beyond mental illness that is both rich and satisfying.

Abbreviations

Community managed/non-government organisations

Clinical rehabilitation

Evidence-based interventions

General Practitioner

International classification of diseases

International classification of functioning, disability and health

Integrated recovery-oriented model

Intermediate stay mental health unit

Mental health

Mental health recovery star

New South Wales

  • Serious mental illness

World health organisation

Andresen R, Oades L, Caputi P. The experience of recovery from schizophrenia: towards an empirically validated stage model. Aust N Z J Psychiatry. 2003;37:586–94.

Article   PubMed   Google Scholar  

Bellack AS. Scientific and consumer models of recovery in schizophrenia: Concordance, contrasts, and implications. Schizophr Bull. 2006;32:432–42.

Article   PubMed   PubMed Central   Google Scholar  

Davidson L. The recovery movement: Implications for mental health care and enabling people to participate fully in life. Health Aff. 2016;35:1091–7.

Article   Google Scholar  

Jose D, Ramachandra, Lalitha K, Gandhi S, Desai G, Nagarajaiah. Consumer perspectives on the concept of recovery in schizophrenia: A systematic review. Asian J Psychiatry. 2015;14:13–8.

Liberman RP, Kopelowicz A. Recovery from schizophrenia: A concept in search of research. Psychiatr Serv. 2005;56:735–42.

Jacobson N, Curtis L. Recovery as policy in mental health services: Strategies emerging from the states. Psychiatr Rehabil J. 2000;23:333–41.

Dickerson FB. Commentary: disquieting aspects of the recovery paradigm. Psychiatr Serv. 2006;57:647.

Johnson S. Misuse and abuse of ‘recovery’ by the psychiatric system. Ment Health Today. 2005;37.

Slade M, Adams N, O’Hagan M. Recovery: past progress and future challenges. Int Rev Psychiatry. 2012;24:1–4.

Drake RE, Whitley R. Recovery and severe mental illness: description and analysis. Can J Psychiatry. 2014;59:236–42.

PubMed   PubMed Central   Google Scholar  

Walsh J, Boyle J. Improving acute psychiatric hospital services according to inpatient experiences. A user-led piece of research as a means to empowerment. Issues Ment Health Nurs. 2009;30:31–8.

Bird V, Leamy M, Tew J, Le Boutillier C, Williams J, Slade M. Fit for purpose? Validation of a conceptual framework for personal recovery with current mental health consumers. Aust N Z J Psychiatry. 2014;48:644–53.

Repper J, Perkins R. Social Inclusion and Recovery: A Model for Mental Health Practice. London: Bailliere Tindall; 2006.

Google Scholar  

Matheson SL, Shepherd AM, Carr VJ. How much do we know about schizophrenia and how well do we know it? Evidence from the Schizophrenia Library. Psychol Med. 2014;44:3387–405.

Article   CAS   PubMed   Google Scholar  

Lieberman JA, Drake RE, Sederer LI, Belger A, Keefe R, Perkins D, Stroup S. Science and recovery in schizophrenia. Psychiatr Serv. 2008;59:487–96.

Mueser KT, Deavers F, Penn DL, Cassisi JE. Psychosocial treatments for schizophrenia. Annu Rev Clin Psychol. 2013;9:465–97.

Anthony WA, Farkas M. Primer on the psychiatric rehabilitation process. Boston: Boston University. Center for Psychiatric Rehabilitation; 2009.

Meltzer HY. Cognitive factors in schizophrenia: causes, impact, and treatment. CNS Spectr. 2004;9:15–24.

Wykes T, Drake RE. Rehabilitative therapies. In: Lieberman JA, Murray RM, editors. Comprehensive Care of Schizophrenia. New York: Oxford University Press; 2012. p. 182–98.

Kuipers E, Yesufu-Udechuku A, Taylor C, Kendall T. Management of psychosis and schizophrenia in adults: summary of updated NICE guidance. BMJ. 2014;348.

Harrow M, Jobe TH, Faull RN. Does treatment of schizophrenia with antipsychotic medications eliminate or reduce psychosis? A 20-year multi-follow-up study. Psychol Med. 2014;44:3007–16.

Deacon BJ. The biomedical model of mental disorder: A critical analysis of its validity, utility, and effects on psychotherapy research. Clin Psychol Rev. 2013;33:846–61.

Velligan D, Gonzalez J. Rehabilitation and recovery in schizophrenia. Psychiatr Clin North Am. 2007;30:535–48.

World-Bank. The World Bank annual report 1996. Washington: World-Bank; 1996.

Book   Google Scholar  

World-Health-Organisation. The Global Burden of Disease, 2004 update. Geneva: World-Health-Organisation; 2008.

Morgan VA, Waterreus A, Jablensky A, Mackinnon A, McGrath JJ, Carr V, Bush R, Castle D, Cohen M, Harvey C, Galletly C, Stain HJ, Neil AL, McGorry P, Hocking B, Shah S, Saw S. People living with psychotic illness in 2010. The second Australian national survey of psychosis. Aust N Z J Psychiatry. 2012;46:735–52.

Neil AL, Carr VJ, Mihalopoulos C, Mackinnon A, Morgan VA. Costs of psychosis in 2010: Findings from the second Australian National Survey of Psychosis. Aust N Z J Psychiatry. 2014;48:169–82.

Neil AL, Carr VJ, Mihalopoulos C, Mackinnon A, Lewin TJ, Morgan VA. What difference a decade? The costs of psychosis in Australia in 2000 and 2010: Comparative results from the first and second Australian national surveys of psychosis. Aust N Z J Psychiatry. 2014;48:237–48.

Commonwealth of Australia. A national framework for recovery-oriented mental health services: Guide for practitioners and providers. Canberra: Department of Health and Ageing, CoA; 2013.

Commonwealth of Australia. A national framework for recovery-oriented mental health services: Policy and theory. Canberra: Department of Health and Ageing, CoA; 2013.

Commonwealth of Australia. National Standards for Mental Health Services. Canberra: CoA; 2010.

Commonwealth of Australia. Development of the Australian Mental Health Care Classification - public consultation paper 1. Canberra: Independent Hospital Pricing Authority, CoA; 2015.

Commonwealth of Australia. Consultation paper on the Pricing Framework for Australian Public Hospital Services 2016–17. Canberra: Independent Hospital Pricing Authority, CoA; 2015.

NSW Department of Health. A New Direction for Mental Health. Sydney: NSW Health; 2006.

Frost BG, Turrell M, Sly KA, Lewin TJ, Conrad AM, Tirupati S, Petrovic K, Rajkumar S, Johnston S: Implementation of a recovery-oriented model in a sub-acute Intermediate Stay Mental Health Unit (ISMHU). BMC Health Services Research (In Press) 2016.

Kidd SA, McKenzie KJ, Virdee G. Mental health reform at a systems level: Widening the lens on recovery-oriented care. Can J Psychiatry. 2014;59:243–9.

Le Boutillier C, Leamy M, Bird VJ, Davidson L, Williams J, Slade M. What does recovery mean in practice? A qualitative analysis of international recovery-oriented practice guidance. Psychiatr Serv. 2011;62:1470–6.

Slade M, Amering M, Farkas M, Hamilton B, O’Hagan M, Panther G, Perkins R, Shepherd G, Tse S, Whitley R. Uses and abuses of recovery: implementing recovery-oriented practices in mental health systems. World Psychiatry. 2014;13:12–20.

Waldemar AK, Arnfred SM, Petersen L, Korsbek L. Recovery-oriented practice in mental health inpatient settings: A literature review. Psychiatr Serv. 2016;67:596–602.

Harding CM, Brooks GW, Ashikaga T, Strauss JS, Breier A. The Vermont longitudinal study of persons with severe mental illness, II: Long-term outcome of subjects who retrospectively met DSM-III criteria for schizophrenia. Am J Psychiatry. 1987;144:727–35.

McGlashan TH. A selective review of recent North American long-term followup studies of schizophrenia. Schizophr Bull. 1988;14:515–42.

Crosse C, Hocking B. Discussion Paper: Social Rehabilitation: What are the Issues? Canberra: Department of Veteran Affairs; 2004.

Harrow M, Grossman LS, Jobe TH, Herbener ES. Do patients with schizophrenia ever show periods of recovery? A 15-year multi-follow-up study. Schizophr Bull. 2005;31:723–34.

Mojtabai R, Fochtmann L, Chang SW, Kotov R, Craig TJ, Bromet E. Unmet need for mental health care in schizophrenia: an overview of literature and new data from a first-admission study. Schizophr Bull. 2009;35:679–95.

Torres-González F, Ibanez-Casas I, Saldivia S, Ballester D, Grandón P, Moreno-Küstner B, Xavier M, Gómez-Beneyto M. Unmet needs in the management of schizophrenia. Neuropsychiatr DisTreat. 2014;10:97–110.

Snyder CR, Ritschel LA, Rand KL, Berg CJ. Balancing psychological assessments: Including strengths and hope in client reports. J Clin Psychol. 2006;62:33–46.

Victorian Department of Health. Framework for recovery-oriented practice. Melbourne: State of Victoria, Department of Health; 2011.

Copic V, Deane FP, Crowe TP, Oades LG. Hope, meaning and responsibility across stages of recovery for individuals living with an enduring mental illness. Aust J Rehabil Couns. 2011;17:61–73.

Weinberg CM. Hope, meaning, and purpose: Making recovery possible. Aust J Rehabil Couns. 2013;36:124–5.

World-Health-Organisation. World Report on Disability. Geneva: World-Health-Organisation; 2011.

Kidd S, Kenny A, McKinstry C. The meaning of recovery in a regional mental health service: an action research study. J Adv Nurs. 2015;71:181–92.

Chen SP, Krupa T, Lysaght R, McCay E, Piat M. The development of recovery competencies for in-patient mental health providers working with people with serious mental illness. Adm Policy Ment Health. 2013;40:96–116.

McKenna B, Furness T, Dhital D, Ennis G, Houghton J, Lupson C, Toomey N. Recovery-oriented care in acute inpatient mental health settings: an exploratory study. Issues Ment Health Nurs. 2014;35:526–32.

Slade M, Williams J, Bird V, Leamy M, Le Boutillier C. Recovery grows up. J Ment Health. 2012;21:99–104.

De Vecchi N, Kenny A, Kidd S. Stakeholder views on a recovery-oriented psychiatric rehabilitation art therapy program in a rural Australian mental health service: a qualitative description. Int J Ment Health Syst. 2015;9.

Slade M, Longden E. Empirical evidence about recovery and mental health. BMC Psychiatry. 2015;15:285.

Reed GM, Leonardi M, Ayuso-Mateos JL, Materzanini A, Castronuovo D, Manara A, Martinez-Aran A, Starace F, Ajovalasit D. Implementing the ICF in a psychiatric rehabilitation setting for people with serious mental illness in the Lombardy region of Italy. Disabil Rehabil. 2009;31:S170–3.

Kilian R, Lauber C, Kalkan R, Dorn W, Rossler W, Wiersma D, Van Buschbach JT, Fioritti A, Tomov T, Catty J, Burns T, Becker T. The relationships between employment, clinical status, and psychiatric hospitalisation in patients with schizophrenia receiving either IPS or a conventional vocational rehabilitation programme. Soc Psychiatry Psychiatr Epidemiol. 2012;47:1381–9.

Perkins R, Slade M. Recovery in England: Transforming statutory services? Int Rev Psychiatry. 2012;24:29–39.

Hopper K. Rethinking social recovery in schizophrenia: What a capabilities approach might offer. Soc Sci Med. 2007;65:868–79.

Rodgers ML, Norell DM, Roll JM, Dyck DG. An overview of mental health recovery. Primary Psychiatry. 2007;14:76–85.

Thornicroft G, Tansella M. Components of a modern mental health service: a pragmatic balance of community and hospital care: Overview of systematic evidence. Br J Psychiatry. 2004;185:283–90.

Flannery F, Adams D, O’Connor N. A community mental health service delivery model: Integrating the evidence base within existing clinical models. Australas Psychiatry. 2011;19:49–55.

Leamy M, Bird V, Le Boutillier C, Williams J, Slade M. Conceptual framework for personal recovery in mental health: Systematic review and narrative synthesis. Br J Psychiatry. 2011;199:445–52.

Smith-Merry J, Freeman R, Sturdy S. Implementing recovery: an analysis of the key technologies in Scotland. Int J Ment Health Syst. 2011;5.

Mueser KT, Corrigan PW, Hilton DW, Tanzman B, Schaub A, Gingerich S, Essock SM, Tarrier N, Morey B, Vogel-Scibilia S, Herz MI. Illness management and recovery: A review of the research. Psychiatr Serv. 2002;53:1272–84.

Slade M, Bird V, Le Boutillier C, Williams J, McCrone P, Leamy M. REFOCUS Trial: protocol for a cluster randomised controlled trial of a pro-recovery intervention within community based mental health teams. BMC Psychiatry. 2011;11.

Slade M, Bird V, Clarke E, Le Boutillier C, McCrone P, Macpherson R, Pesola F, Wallace G, Williams J, Leamy M. Supporting recovery in patients with psychosis through care by community-based adult mental health teams (REFOCUS): a multisite, cluster, randomised, controlled trial. Lancet Psychiatry. 2015;2:503–14.

Tse S, Tsoi EWS, Hamilton B, O’Hagan M, Shepherd G, Slade M, Whitley R, Petrakis M. Uses of strength-based interventions for people with serious mental illness: A critical review. Int J Soc Psychiatry. 2016;62:281–91.

Bond GR, Drake RE, Becker DR, Noel VA. Sustaining Individual Placement and Support (IPS) services: the IPS learning community. World Psychiatry. 2016;15:81–3.

Petrakis M, Brophy L, Lewis J, Stylianou M, Scott M, Cocks N, Buckley L, Halloran K. Consumer measures and research co-production: a pilot study evaluating the recovery orientation of a mental health program collaboration. Asia Pac J Soc Work Dev. 2014;24:94–108.

Gilburt H, Slade M, Bird V, Oduola S, Craig TK. Promoting recovery-oriented practice in mental health services: a quasi-experimental mixed-methods study. BMC Psychiatry. 2013;13:167.

Horsfall J, Cleary M, Hunt GE. Acute inpatient units in a comprehensive (integrated) mental health system: a review of the literature. Issues Ment Health Nurs. 2010;31:273–8.

Le Boutillier C, Chevalier A, Lawrence V, Leamy M, Bird VJ, Macpherson R, Williams J, Slade M. Staff understanding of recovery-orientated mental health practice: a systematic review and narrative synthesis. Implement Sci. 2015;10.

Jacobson N, Greenley D. What is recovery? A conceptual model and explication. Psychiatr Serv. 2001;52:482–5.

Le Boutillier C, Slade M, Lawrence V, Bird VJ, Chandler R, Farkas M, Harding C, Larsen J, Oades LG, Roberts G, Shepherd G, Thornicroft G, Williams J, Leamy M. Competing priorities: Staff perspectives on supporting recovery. Adm Policy Ment Health Ment Health Serv Res. 2015;42:429–38.

Anthony WA, Farkas M. The essential guide to psychiatric rehabilitation practice. Boston: Boston University; 2012.

Wolwer W, Frommann N. Social-cognitive remediation in schizophrenia: generalization of effects of the Training of Affect Recognition (TAR). Schizophr Bull. 2011;37(2):S63–70.

Velligan D, Ritch J, Sui D, DiCocco M, Huntzinger C. Frontal Systems Behavior Scale in schizophrenia: relationships with psychiatric symptomatology, cognition and adaptive function. Psychiatry Res. 2002;113:227–36.

Roder V, Mueller DR, Schmidt SJ. Effectiveness of Integrated Psychological Therapy (IPT) for schizophrenia patients: A research update. Schizophr Bull. 2011;37:S71–9.

Verhaeghe N, De Maeseneer J, Maes L, Van Heeringen C, Annemans L. Effectiveness and cost-effectiveness of lifestyle interventions on physical activity and eating habits in persons with severe mental disorders: a systematic review. Int J Behav Nutr Phys Act. 2011;8:28.

Killackey E, Waghorn G. The challenge of integrating employment services with public mental health services in Australia: Progress at the first demonstration site. Psychiatr Rehabil J. 2008;32:63–6.

Lundahl B, Burke BL. The effectiveness and applicability of motivational interviewing: a practice-friendly review of four meta-analyses. J Clin Psychol. 2009;65:1232–45.

Miller WR, Rose GS. Toward a theory of motivational interviewing. Am Psychol. 2009;64:527–37.

Kern RS, Green MF, Mintz J, Liberman RP. Does ‘errorless learning’ compensate for neurocognitive impairments in the work rehabilitation of persons with schizophrenia? Psychol Med. 2003;33:433–42.

Lyman DR, Kurtz MM, Farkas M, George P, Dougherty RH, Daniels AS, Ghose SS, Delphin-Rittmon ME. Skill building: Assessing the evidence. Psychiatr Serv. 2014;65:727–38.

Pilling S, Bebbington P, Kuipers E, Garety P, Geddes J, Orbach G, Morgan C. Psychological treatments in schizophrenia: I. Meta-analysis of family intervention and cognitive behaviour therapy. Psychol Med. 2002;32:763–82.

CAS   PubMed   Google Scholar  

Borg M, Kristiansen K. Recovery-oriented professionals: Helping relationships in mental health services. J Ment Health. 2004;13:493–505.

World-Health-Organisation. The International Classification of Functioning, Disability, and Health. Geneva: WHO; 2001.

MacKeith J, Burns S. Mental Health Recovery Star; User Guide. Secondth ed. London: Triangle Consulting and Mental Health Providers Forum; 2010.

Parker S, Dark F, Newman E, Korman N, Meurk C, Siskind D, Harris M. Longitudinal comparative evaluation of the equivalence of an integrated peer-support and clinical staffing model for residential mental health rehabilitation: a mixed methods protocol incorporating multiple stakeholder perspectives. BMC Psychiatry. 2016;16:1–21.

Burgess P, Pirkis J, Coombs T, Rosen A. Assessing the value of existing recovery measures for routine use in Australian mental health services. Aust N Z J Psychiatry. 2011;45:267–80.

Andresen R, Caputi P, Oades LG. Do clinical outcome measures assess consumer-defined recovery? Psychiatry Res. 2010;177:309–17.

Shanks V, Williams J, Leamy M, Bird VJ, Le Boutillier C, Slade M. Measures of personal recovery: A systematic review. Psychiatr Serv. 2013;64:974–80.

Chester P, Ehrlich C, Warburton L, Baker D, Kendall E, Crompton D. What is the work of recovery oriented practice? A systematic literature review. Int J Ment Health Nurs. 2016;25:270–85.

Tsai J, Salyers MP. Recovery orientation in hospital and community settings. J Behav Health Serv Res. 2010;37:385–99.

Williams J, Leamy M, Bird V, Harding C, Larsen J, Le Boutillier C, Oades L, Slade M. Measures of the recovery orientation of mental health services: systematic review. Soc Psychiatry Psychiatr Epidemiol. 2012;47:1827–35.

Whitley R, Drake RE. Recovery: a dimensional approach. Psychiatr Serv. 2010;61:1248–50.

Download references

Acknowledgements

Thank you to the Executive of Hunter New England MH Services for their support and to Psychiatric Rehabilitation Services’ staff and clients for their contributions to model development and evaluation. We would also like to thank the acute and community MH clinicians who completed the initial staff surveys.

No external funding was received for this project.

Availability of data and materials

While data from this project will not be made publicly available, we will undertake reasonable requests for additional analyses.

Authors’ contributions

BGF: initiated the project and model development, contributed to project design, implementation, interpretation, and drafting of the manuscript. ST, SJ and MT: contributed to model development, project design, implementation, interpretation, and manuscript revision. KAS and TJL: contributed to project design, data collection, statistical analysis and interpretation, manuscript preparation and revision. AMC: contributed to project design, interpretation, and manuscript revision. All authors read and approved the final manuscript.

Competing interests

The authors report no conflicts of interest, and that they alone are responsible for the paper’s content.

Consent for publication

Not applicable, as the data analyses undertaken do not report any individual’s data.

Ethics approval and consent to participate

This project received confirmation from Hunter New England Human Research Ethics Committee that it was exempt from formal review (letter dated October 24th 2013), being viewed as part of an internal, low risk, service evaluation.

Author information

Authors and affiliations.

School of Psychology, Faculty of Science and Technology, University of Newcastle, Callaghan, NSW, 2308, Australia

Barry G. Frost

Centre for Brain and Mental Health Research, Hunter New England Mental Health and the University of Newcastle, Callaghan, NSW, 2308, Australia

Barry G. Frost, Terry J. Lewin, Ketrina A. Sly & Agatha M. Conrad

Hunter New England Mental Health, Newcastle, NSW, 2300, Australia

Srinivasan Tirupati, Suzanne Johnston, Megan Turrell, Terry J. Lewin, Ketrina A. Sly & Agatha M. Conrad

School of Medicine and Public Health, Faculty of Health and Medicine, University of Newcastle, Callaghan, NSW, 2308, Australia

Srinivasan Tirupati, Terry J. Lewin, Ketrina A. Sly & Agatha M. Conrad

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Terry J. Lewin .

Additional file

Additional file 1:.

Perceived impacts on and relevance of ‘recovery domains’: Responses from a recent survey of MH clinicians ( N  = 251). (DOCX 17 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License ( http://creativecommons.org/licenses/by/4.0/ ), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Cite this article.

Frost, B.G., Tirupati, S., Johnston, S. et al. An Integrated Recovery-oriented Model (IRM) for mental health services: evolution and challenges. BMC Psychiatry 17 , 22 (2017). https://doi.org/10.1186/s12888-016-1164-3

Download citation

Received : 18 August 2016

Accepted : 08 December 2016

Published : 17 January 2017

DOI : https://doi.org/10.1186/s12888-016-1164-3

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Evidence-based psychosocial interventions
  • Mental health services
  • Recovery-oriented
  • Rehabilitation

BMC Psychiatry

ISSN: 1471-244X

recovery model research paper

  • - Google Chrome

Intended for healthcare professionals

  • Access provided by Google Indexer
  • My email alerts
  • BMA member login
  • Username * Password * Forgot your log in details? Need to activate BMA Member Log In Log in via OpenAthens Log in via your institution

Home

Search form

  • Advanced search
  • Search responses
  • Search blogs
  • News & Views
  • The recovery model and...

The recovery model and patient generated outcome measures

  • Related content
  • Peer review
  • Heidi J Hales , psychiatrist
  • North West London Forensic Child and Adolescent Mental Health Service, St Bernard’s Hospital, London UB1 3EU, UK
  • heidi.hales{at}nhs.net

Robertson discusses the notions of cure and wellness. 1 In mental health, the vision of healing as “thriving within one’s life as it is” is embedded in the recovery model. We always aim to reduce distressing symptoms, but labelling people as unwell and incurable can cause its own iatrogenic harm. Those living in more accepting, inclusive, and caring societies experience less mental illness than those in our society (which is focused on health, wealth, independence, and personal ambition) because they are accepted as they are.

As Robertson says, the start is not for us to assume what a good outcome is but to ask our client what they hope for and then work with them to achieve that. This is true collaborative coproduction of care, which builds trust that can help our clients through stressful and challenging times, relapses, drug side effects, and hospital admissions. If we acknowledge that life has its ups and downs and that health professionals can provide support at points of crisis, caring not judging, supporting not enforcing, I think we will be seen as more accessible in the next time of crisis.

This model of care and understanding not only improves our clinical practice but should also be embedded in clinical trials—let’s get patient generated outcome measures embedded in all our trials, ensuring that we are truly meeting patients’ goals and not those we assume they have. Then we can truly assess whether “treatment” offers good outcomes.

Competing interests: None declared.

  • Robertson AD

recovery model research paper

Change Password

Your password must have 6 characters or more:.

  • a lower case character, 
  • an upper case character, 
  • a special character 

Password Changed Successfully

Your password has been changed

Create your account

Forget yout password.

Enter your email address below and we will send you the reset instructions

If the address matches an existing account you will receive an email with instructions to reset your password

Forgot your Username?

Enter your email address below and we will send you your username

If the address matches an existing account you will receive an email with instructions to retrieve your username

Psychiatry Online

  • April 01, 2024 | VOL. 75, NO. 4 CURRENT ISSUE pp.307-398
  • March 01, 2024 | VOL. 75, NO. 3 pp.203-304
  • February 01, 2024 | VOL. 75, NO. 2 pp.107-201
  • January 01, 2024 | VOL. 75, NO. 1 pp.1-71

The American Psychiatric Association (APA) has updated its Privacy Policy and Terms of Use , including with new information specifically addressed to individuals in the European Economic Area. As described in the Privacy Policy and Terms of Use, this website utilizes cookies, including for the purpose of offering an optimal online experience and services tailored to your preferences.

Please read the entire Privacy Policy and Terms of Use. By closing this message, browsing this website, continuing the navigation, or otherwise continuing to use the APA's websites, you confirm that you understand and accept the terms of the Privacy Policy and Terms of Use, including the utilization of cookies.

What Is Recovery? A Conceptual Model and Explication

  • Nora Jacobson , Ph.D. , and
  • Dianne Greenley , M.S.W., J.D.

Search for more papers by this author

This paper describes a conceptual model of recovery from mental illness developed to aid the state of Wisconsin in moving toward its goal of developing a "recovery-oriented" mental health system. In the model, recovery refers to both internal conditions experienced by persons who describe themselves as being in recovery—hope, healing, empowerment, and connection—and external conditions that facilitate recovery—implementation of the principle of human rights, a positive culture of healing, and recovery-oriented services. The aim of the model is to link the abstract concepts that define recovery with specific strategies that systems, agencies, and individuals can use to facilitate it.

The current notion of recovery from mental illness dates back to the 1980s, with the publication of a major study that demonstrated that the course of severe mental illness was not an inevitable deterioration ( 1 ) and of several first-person accounts by consumers who described their experiences with a diagnosed mental illness and how they had managed to emerge intact or recover ( 2 , 3 , 4 ). Mental health professionals drew on such accounts to formulate theoretical and practical models of recovery that could be adapted for use in psychosocial rehabilitation and other mental health services ( 5 , 6 , 7 , 8 , 9 ). In the 1990s, as states were faced with the challenge of reconfiguring their publicly funded mental health services according to the principles of managed care ( 10 ), recovery became a tool for guiding system reform in both policy and practice ( 11 ).

The use of the term "recovery" in these different arenas—outcomes research, personal narrative, services design and provision, and system reform—has led to confusion. Recovery is variously described as something that individuals experience, that services promote, and that systems facilitate, yet the specifics of exactly what is to be experienced, promoted, or facilitated—and how—are often not well understood either by the consumers who are expected to recover or by the professionals and policy makers who are expected to help them.

The conceptual model of recovery described in this paper was designed for the purposes of education and self-assessment ( 12 ). It was developed by the first author under the aegis of the State of Wisconsin's recovery implementation task force, an advisory body composed of consumers, providers, advocates, and policy makers that has been charged with moving the state toward its goal of developing a "recovery-oriented" mental health system ( 13 ). The model aims to link the abstract concepts that define recovery with the specific strategies that systems, agencies, and individuals can use to facilitate it.

In our model, the word recovery refers both to internal conditions—the attitudes, experiences, and processes of change of individuals who are recovering—and external conditions—the circumstances, events, policies, and practices that may facilitate recovery. Together, internal and external conditions produce the process called recovery. These conditions have a reciprocal effect, and the process of recovery, once realized, can itself become a factor that further transforms both internal and external conditions.

Internal conditions

An analysis of numerous accounts by consumers of mental health services who describe themselves as "being in recovery" or "on a journey of recovery" suggests that the key conditions in this process are hope, healing, empowerment, and connection.

Hope. The hope that leads to recovery is, at its most basic level, the individual's belief that recovery is possible. The attitudinal components of hope are recognizing and accepting that there is a problem, committing to change, focusing on strengths rather than on weaknesses or the possibility of failure, looking forward rather than ruminating on the past, celebrating small steps rather than expecting seismic shifts in a short time, reordering priorities, and cultivating optimism. Gaining hope has about it something of the transcendent. "A tiny, fragile spark of hope appeared and promised that there could be something more than all of this darkness.… This is the mystery. This is the grace.… All of the polemic and technology of psychiatry, psychology, social work, and science cannot account for this phenomenon of hope. But those of us who have recovered know that this grace is real. We lived it. It is our shared secret" ( 3 ).

The source of this grace is different for each individual. For one it will be the entity he or she knows as God. For another, it might be a spiritual connection with nature. Individuals not drawn to spirituality may find their grace in other sources, such as making art or contemplating philosophical issues. Hope sustains, even during periods of relapse. It creates its own possibilities. Hope is a frame of mind that colors every perception. By expanding the realm of the possible, hope lays the groundwork for healing to begin.

Healing. Consumers and professionals who accept the dictionary definition of recovery—to regain normal health, poise, or status—may resist the very possibility of recovery because they see it as an unrealistic expectation. However, it is important to remember that recovery is not synonymous with cure. Recovery is distinguished both by its endpoint—which is not necessarily a return to "normal" health and functioning—and by its emphasis on the individual's active participation in self-help activities. The concept of recovery is better captured by the notion of healing, a process that has two main components: defining a self apart from illness, and control.

As Estroff ( 14 ) has noted, people who have psychiatric disabilities often find that they lose their "selves" inside mental illness. Recovery is in part the process of "recovering" the self by reconceptualizing illness as only a part of the self, not as a definition of the whole. As consumers reconnect with their selves, they begin to experience a sense of self-esteem and self-respect that allows them to confront and overcome the stigma against persons with mental illness that they may have internalized, thus allowing further connection with the self.

The second healing process is control—that is, finding ways to relieve the symptoms of the illness or reduce the social and psychological effects of stress. For some consumers, medication is a successful strategy for effecting control. Another strategy is learning to reduce the occurrence and severity of symptoms and the effects of stress through self-care practices, such as adopting a wellness lifestyle or using symptom monitoring and response techniques ( 15 , 16 )

The word "control" has a double meaning. In one sense it refers to the outcome of managing symptoms or stress. The second meaning, however, refers to the locus of control, or who has control. In recovery it is the consumer who has taken control, who has become an active agent in his or her own life. Control is an important factor in the next internal condition, empowerment.

Empowerment. In its simplest sense, empowerment may be understood as a corrective for the lack of control, sense of helplessness, and dependency that many consumers develop after long-term interactions with the mental health system. A sense of empowerment emerges from inside one's self—although it may be facilitated by external conditions—and it has three components. The first is autonomy, or the ability to act as an independent agent. The tools needed to act autonomously include knowledge, self-confidence, and the availability of meaningful choices. The second is courage—a willingness to take risks, to speak in one's own voice, and to step outside of safe routines. The third is responsibility, a concept that speaks to the consumer's obligations.

In the recovery model, the aim is to have consumers assume more and more responsibility for themselves. Their particular responsibilities include developing goals, working with providers and others—for example, family and friends—to make plans for reaching these goals, taking on decision-making tasks, and engaging in self-care. In addition, responsibility is a factor in making choices and taking risks; full empowerment requires that consumers live with the consequences of their choices.

Connection. Recovery is a profoundly social process. As consumers' accounts make clear, much of what is being recovered is a way of being in the company of others. The internal condition called connection captures the aspect of recovery that has to do with rejoining the social world—what some have called "getting a life." The ability to forge connections with others is both a result of hope, healing, and empowerment and a way to make these internal conditions possible.

To connect is to find roles to play in the world. These roles may involve activities, relationship status, or occupation. Many consumers report that the most powerful form of connection is helping others who are also living with mental illness. For some consumers, this means becoming a mental health provider or advocate; for others, it means bearing witness, or telling their own stories in public arenas. In all of these capacities, consumers increase the general understanding of what it is like to live with a mental illness. They find ways to validate and reconcile their own experiences, and by standing as living exemplars of the possibility of recovery, they serve as role models for others.

In yet another sense, connection is the bridge between internal and external conditions, allowing reciprocal action between the two.

External conditions

The external conditions that define recovery are human rights, "a positive culture of healing" ( 17 ), and recovery-oriented services. On the surface, these three conditions seem quite different. Human rights denotes a broad, societal condition; a positive culture of healing refers to the cultural milieu in which services are offered; and recovery-oriented services are the actual services provided. It is important to recognize, however, that these three conditions are simply different foci viewed through the same lens. That is, implementation of the principles of human rights in an organization results in a positive culture of healing, and recovery-oriented services are services that emerge from such a culture.

Human rights. In its broadest sense, a human rights agenda lays out a vision of a society in which power and resources are distributed equitably. When applied to mental illness, human rights emphasizes reducing and then eliminating stigma and discrimination against persons with psychiatric disabilities; promoting and protecting the rights of persons in the service system; providing equal opportunities for consumers in education, employment, and housing; and ensuring that consumers have access to needed resources, including those necessary for sustaining life (adequate food and shelter) as well as the social and health services that can aid recovery (physical, dental, and mental health services; job training; supported housing; and employment programs).

This human rights agenda allows for different perspectives and different types of activism. It can be used to advocate for the reduction and ultimately the elimination of involuntary commitment and other forced treatment, which many view as violations of human rights, or it can be used to campaign for parity legislation and universal health coverage. In this formulation, equal opportunity is promoted through expanded access to care and services that might otherwise be restricted by poverty, stigma, or the law itself.

A positive culture of healing. Fisher ( 17 ) has written of the need to "build a coherent social faith and order" as a way to promote recovery. He described this new order as "a positive culture of healing… a culture of inclusion, caring, cooperation, dreaming, humility, empowerment, hope, humor, dignity, respect, trust, and love." When applied to the culture of a human services organization, this vision of a positive culture of healing begins with an environment characterized by tolerance, listening, empathy, compassion, respect, safety, trust, diversity, and cultural competence. A healing culture is oriented toward human rights for all individuals and groups. Consumers' rights are incorporated into all decisions, and informed consent is part of the bedrock of daily practice.

In a positive culture of healing, professionals as well as consumers are empowered and engaged. For providers, empowerment means first believing that they can make a difference and then making a commitment to changing the way they conceptualize the course of mental illness and the way they practice. Providers must embrace the belief that every consumer can achieve hope, healing, empowerment, and connection, no matter what his or her current status. This belief must lead them to focus on the person, not the illness, and on his or her strengths and goals.

A key component of a positive culture of healing is the development of collaborative relationships between consumers and providers. In contrast to a hierarchical model of service provision, the collaborative model allows consumers and providers to work together to plan, negotiate, and make decisions about the services and activities the consumer will use to support his or her recovery. Collaboration implies that the consumer is an active participant, that he or she is presented with a range of options and given the opportunity to choose from among them, and that providers allow the consumer to take some risks with these choices. Consumers have the opportunity to make choices other than those the provider might have made for them.

Finally, a true collaborative relationship is one in which both consumer and provider come to see each other as human beings. For providers, this means learning to see beyond the diagnostic—or racial, ethnic, and socioeconomic—categories they have been trained to use and rethinking "boundary issues" so they can allow themselves to relate to consumers on a human level.

Recovery-oriented services. The Boston University Center for Psychiatric Rehabilitation has developed a model for designing recovery-oriented services ( 5 , 7 ). The model delineates four major consequences of severe mental illness—impairment, dysfunction, disability, and disadvantage. Recovery-oriented services address the range of these features and include services directed at symptom relief, crisis intervention, case management, rehabilitation, enrichment, rights protection, basic support, and self-help.

A second model, developed by the Ohio Department of Mental Health ( 18 ), describes the best practices to be implemented by consumers, clinicians, and community supports at four different stages of the mental health recovery process. The practices encompass clinical care, peer and family support, work, power and control, stigma, community involvement, access to resources, and education. A third model ( 19 ) offers practice guidance within "a framework for designing, implementing, and evaluating behavior healthcare services that facilitate individual recovery and personal outcomes." Using the overarching metaphor of "a healing culture," this model addresses such issues as language, dignity and respect, empowerment and personal responsibility, consumer and family involvement, challenging stigma and discrimination, reflective practice and continuous improvement, cultural sensitivity and safety, and spirituality and personal meaning.

Each of these models integrates services provided by professionals, services provided by consumers, and services provided in collaboration. Services provided by professionals include medication, psychiatric rehabilitation, and traditional support services such as therapy and case management. The recovery orientation in these services lies in the attitudes of the professionals who provide them. For example, decisions about medication are worked out in a partnership between the provider and the consumer, rather than being dictated by the provider.

Consumer-run services are planned, implemented, and provided by consumers for consumers. Examples include advocacy, peer support programs, hospitalization alternatives, hotlines or "warm lines," and programming that provides opportunities for role modeling and mentoring.

Collaborative services are provided by and for both consumers and professionals as well as family members, friends, and members of the larger community and emphasize their diverse but complementary strengths. Examples include recovery education and training, clubhouse organizations, crisis planning, the development of recovery and treatment plans, community integration, and consumer rights education.

Although many of these services may sound similar to services currently being offered in many mental health systems, it is important to recognize that no service is recovery-oriented unless it incorporates the attitude that recovery is possible and has the goal of promoting hope, healing, empowerment, and connection.

Conclusions

The reciprocal relationship between the internal and the external conditions of recovery has been implicit in the presentation of this model. For example, reducing social stigma will help reduce the internalized stigma that restricts the ability of some consumers to define a self apart from their diagnosis. Access to appropriate mental health services, including education, will provide consumers with the knowledge, skills, and strategies that can help them relieve symptoms and control the effects of stress. Collaborative relationships between providers and consumers will empower both parties, allowing meaningful power sharing and a more mutual assumption of responsibility. Peer support provides opportunities for bearing witness, a practice that allows the speaker and the listeners to establish new connections and validates the idea that recovery is possible.

Similarly, as more consumers recover, and as recovery becomes more firmly entrenched in policy and practice, internal and external conditions will be transformed. More consumers in recovery will provide more models of what hope, healing, empowerment, and connection might look like. More experience with and evaluation of recovery-oriented services will determine which of these services has the greatest influence on recovery.

Acknowledgment

Dr. Jacobson's work on the conceptual model and on this paper was supported by the Wisconsin Coalition for Advocacy.

Dr. Jacobson is an associate scientist with the University of Wisconsin School of Nursing in Madison, K6/316 Clinical Science Center, 600 Highland Avenue, Madison, Wisconsin 53792 (e-mail, [email protected] ). Ms. Greenley is a supervising attorney at the Wisconsin Coalition for Advocacy and co-principal investigator with the women and mental health study site at the University of Wisconsin School of Social Work.

1. Harding CM, Brooks GW, Ashikaga T, et al: The Vermont longitudinal study of persons with severe mental illness: II. long-term outcome of subjects who retrospectively met DSM-III criteria for schizophrenia. American Journal of Psychiatry 144:727-735, 1987 Link ,  Google Scholar

2. Houghton JF: First-person account: maintaining mental health in a turbulent world. Schizophrenia Bulletin 8:548-552, 1982 Crossref , Medline ,  Google Scholar

3. Deegan PE: Recovery: the lived experience of rehabilitation. Psychosocial Rehabilitation Journal 11(4):11-19, 1988 Google Scholar

4. Unzicker R: On my own: a personal journey through madness and re-emergence. Psychosocial Rehabilitation Journal 13(1):71-77, 1989 Google Scholar

5. Anthony W: Recovery from mental illness: the guiding vision of the mental health service system in the 1990s. Psychosocial Rehabilitation Journal 16(4):11-23, 1993 Google Scholar

6. Hatfield AB, Lefley HP: Surviving Mental Illness: Stress, Coping, and Adaptation. New York, Guilford, 1993 Google Scholar

7. Spaniol L, Koehler M, Hutchinson D: The Recovery Workbook: Practical Coping and Empowerment Strategies for People With Psychiatric Disability. Boston, Boston University Center for Psychiatric Rehabilitation, 1994 Google Scholar

8. Fisher DB: Health care reform based on an empowerment model of recovery by people with psychiatric disabilities. Hospital and Community Psychiatry 45:913-915, 1994 Abstract ,  Google Scholar

9. Davidson L, Strauss JS: Beyond the biopsychosocial model: integrating disorder, health, and recovery. Psychiatry 58:44-55, 1995 Crossref , Medline ,  Google Scholar

10. Mechanic D: Emerging trends in mental health policy and practice. Health Affairs 17(6):82-98, 1998 Google Scholar

11. Jacobson N, Curtis L: Recovery as policy in mental health services: strategies emerging from the states. Psychiatric Rehabilitation Journal 23(4):333-341, 2000 Google Scholar

12. Jacobson N: A Conceptual Model of Recovery. Madison, Wis, Wisconsin Coalition for Advocacy, 2000 Google Scholar

13. The Blue Ribbon Commission on Mental Health: Final Report. Madison, Wis, 1997 Google Scholar

14. Estroff SE: Self, identity, and subjective experiences of schizophrenia: in search of the subject. Schizophrenia Bulletin 15:189-196, 1989 Crossref , Medline ,  Google Scholar

15. Copeland ME: WRAP: Wellness Recovery Action Plan. Brattleboro, Vt, Peach Press, 1997 Google Scholar

16. Crowley K: The Power of Procovery in Healing Mental Illness. San Francisco, Kennedy-Carlisle, 2000 Google Scholar

17. Fishe DB: Towards a positive culture of healing, in The DMH Core Curriculum: Consumer Empowerment and Recovery, Part I. Boston, Commonwealth of Massachusetts Department of Mental Health, 1993 Google Scholar

18. Townsend W, Boyd S, Griffin G, et al: Emerging best practices in mental health recovery. Columbus, Ohio Department of Mental Health, undated Google Scholar

19. Curtis LC: Practice guidance for recovery-oriented behavioral healthcare for adults with serious mental illnesses, in Personal Outcome Measures in Consumer-Directed Behavioral Health. Towson, Md, Council on Quality and Leadership for Persons With Disabilities, 2000 Google Scholar

  • Resilience and Recovery in the Context of Psychological Disorders 29 May 2019 | Journal of Humanistic Psychology, Vol. 64, No. 3
  • Spirituality and Recovery From Severe Mental Disorders: A Systematic Review 8 March 2024 | Journal of Psychosocial Nursing and Mental Health Services, Vol. 18
  • Plant-based recovery from restrictive eating disorder: A qualitative enquiry Appetite, Vol. 194
  • Examining an Alternate Care Pathway for Mental Health and Addiction Prehospital Emergencies in Ontario, Canada: A Critical Analysis 29 January 2024 | International Journal of Environmental Research and Public Health, Vol. 21, No. 2
  • What should inpatient psychological therapies be for? Qualitative views of service users on outcomes 12 October 2023 | Health Expectations, Vol. 27, No. 1
  • The experience of community mental health teams by people with long-term experience of psychosis 4 January 2023 | Psychosis, Vol. 16, No. 1
  • The Poetry of Recovery in Peer Support Workers with Mental Illness: An Interpretative Phenomenological Analysis 5 January 2024 | Healthcare, Vol. 12, No. 2
  • Recovery-orientierte Behandlung und Genesungsbegleitung in der Psychiatrie 15 January 2024 | Der Nervenarzt, Vol. 95, No. 1
  • Hope, resilience and subjective happiness among general population of Paraguay in the post COVID-19 pandemic 7 December 2023 | International Journal of Social Psychiatry, Vol. 29
  • From existing to living: Exploring the meaning of recovery and a sober life after a long duration of a substance use disorder 24 May 2023 | Nordic Studies on Alcohol and Drugs, Vol. 40, No. 6
  • Staff experiences of a novel in‐reach rehabilitation and recovery service for people with profound and enduring mental health needs 3 May 2023 | International Journal of Mental Health Nursing, Vol. 32, No. 5
  • Benefits of the PRISM Shelter-Based Program for Attainment of Stable Housing and Functional Outcomes by People Experiencing Homelessness and Mental Illness: A Quantitative Analysis 20 March 2023 | The Canadian Journal of Psychiatry, Vol. 68, No. 10
  • Developing a consensus of recovery from suicidal ideations and behaviours: A Delphi study with experts by experience 20 September 2023 | PLOS ONE, Vol. 18, No. 9
  • Therapists who foster social identification build stronger therapeutic working alliance and have better client outcomes Comprehensive Psychiatry, Vol. 124
  • Long-term outcomes of the recovery approach in a high-security mental health setting: a 20 year follow-up study 11 May 2023 | Frontiers in Psychiatry, Vol. 14
  • Idiographic patient reported outcome measures (I‐PROMs) for routine outcome monitoring in psychological therapies: Position paper 22 February 2022 | Journal of Clinical Psychology, Vol. 79, No. 3
  • Examining Implementation of Crisis Centers on Police Officer Emergency Hold Petitions 15 February 2023 | Community Mental Health Journal, Vol. 57
  • Grounding co‐writing: An analysis of the theoretical basis of a new approach in mental health care 28 April 2022 | Journal of Psychiatric and Mental Health Nursing, Vol. 30, No. 1
  • Hope and resilience among patients affected by unipolar and bipolar depression 6 January 2023 | International Journal of Social Psychiatry, Vol. 29
  • Per una salute mentale di comunità: la questione dei Servizi e dei modelli PSICOTERAPIA E SCIENZE UMANE, No. 1
  • Adult Attachment and Personal Recovery in Clients With a Psychotic Disorder 31 March 2023 | Schizophrenia Bulletin Open, Vol. 4, No. 1
  • Working Through Relational Trauma: An Exploration of Narratives of Lived Experiences of Trauma and Recovery 15 February 2024
  • The Effect of Self-Stigma on the Hope of Chinese with Mental Illness: The Mediating Role of Family Function 12 December 2022 | Psychiatry, Vol. 6
  • Strength- and recovery-based approaches in forensic mental health in late modernity: Increasingly incorporating a human rights angle? 13 July 2021 | Social Theory & Health, Vol. 20, No. 4
  • A mixed methods study examining perceptions by service-users of their involuntary admission in relation to levels of insight 22 December 2021 | International Journal of Social Psychiatry, Vol. 68, No. 8
  • Older adults' perspectives on rehabilitation and recovery one year after a hip fracture – a qualitative study 14 May 2022 | BMC Geriatrics, Vol. 22, No. 1
  • Assessing a digital peer support self-management intervention for adults with serious mental illness: feasibility, acceptability, and preliminary effectiveness 28 January 2022 | Journal of Mental Health, Vol. 31, No. 6
  • Positive psychology in neurodiversity: An investigation of character strengths in autistic adults in the United Kingdom in a community setting Research in Autism Spectrum Disorders, Vol. 99
  • Shared decision-making interventions for people with mental health conditions 11 November 2022 | Cochrane Database of Systematic Reviews, Vol. 2022, No. 11
  • A Scoping Review of Resilience in Survivors of Human Trafficking 20 January 2021 | Trauma, Violence, & Abuse, Vol. 23, No. 4
  • Guérison autochtone et rétablissement en santé mentale : rapprochement conceptuel et implications pour l’intervention psychosociale Canadian Journal of Community Mental Health, Vol. 41, No. 4
  • Feasibility, Acceptability, and Potential Utility of Peer-supported Ecological Momentary Assessment Among People with Serious Mental Illness: a Pilot Study 4 June 2022 | Psychiatric Quarterly, Vol. 93, No. 3
  • The experience of Hope among Italian undergraduate students 15 September 2020 | Current Psychology, Vol. 41, No. 8
  • Making People Autonomous: A Sociological Analysis of the Uses of Contracts and Projects in the Psychiatric Care Institutions 5 March 2019 | Culture, Medicine, and Psychiatry, Vol. 46, No. 2
  • Models and frameworks of mental health recovery: a scoping review of the available literature 10 May 2022 | Journal of Mental Health, Vol. 8
  • My Space, Your Space, Our Space: Exploring the Potential of Collaborative Group Facilitation Between Therapists and Peer Workers in Mental Health Settings 23 June 2021 | Community Mental Health Journal, Vol. 58, No. 3
  • Engaging peer consultants in mental health services: Narrative research 14 March 2021 | International Journal of Social Psychiatry, Vol. 68, No. 2
  • “For me I see MINE to be a family sickness” – consumers understanding and perception of the etiology of mental illness in community-based residential facilities in Ghana 17 February 2022 | Mental Health, Religion & Culture, Vol. 25, No. 1
  • Incorporações e usos do conceito de recovery no contexto da Reforma Psiquiátrica Brasileira: uma revisão da literatura 1 January 2022 | Trabalho, Educação e Saúde, Vol. 20
  • Post-traumatic Growth and Recovery 21 July 2022
  • Making recovery meaningful for people with intellectual disabilities 21 October 2021 | Journal of Applied Research in Intellectual Disabilities, Vol. 35, No. 1
  • Improvements in Hope, Engagement and Functioning Following a Recovery-Focused Sub-Acute Inpatient Intervention: a Six-Month Evaluation 14 June 2021 | Psychiatric Quarterly, Vol. 92, No. 4
  • Professional perspectives on providing recovery-oriented services in Taiwan: a qualitative study 16 March 2021 | BMC Psychiatry, Vol. 21, No. 1
  • Discrepancy between experience and importance of recovery components in the symptomatic and recovery perceptions of people with severe mental disorders 31 May 2021 | BMC Psychiatry, Vol. 21, No. 1
  • Comprehensive well-being scale: development and validation among Chinese in recovery of mental illness in Hong Kong 13 November 2021 | BMC Psychology, Vol. 9, No. 1
  • “Surviving out of the Ashes” – An exploration of young adult service users' perspectives of mental health recovery 8 February 2021 | Journal of Psychiatric and Mental Health Nursing, Vol. 28, No. 5
  • Impact of psychiatric advance directive facilitation on mental health consumers: empowerment, treatment attitudes and the role of peer support specialists 4 February 2020 | Journal of Mental Health, Vol. 30, No. 5
  • Towards Developing An EMR in Mental Health Care for Children’s Mental Health Development among the Underserved Communities in USA
  • Creative wellbeing. Prototyping an arts-health practice program for mental health recovery 20 May 2021 | Design for Health, Vol. 5, No. 1
  • The Role of Self-efficacy in the Recovery Trajectory of Mental Health Consumers 6 March 2021 | The British Journal of Social Work, Vol. 51, No. 3
  • An Integrative Review of Sensory Approaches in Adult Inpatient Mental Health: Implications for Occupational Therapy in Prison-Based Mental Health Services 5 December 2020 | Occupational Therapy in Mental Health, Vol. 37, No. 2
  • Adult Attachment and Personal, Social, and Symptomatic Recovery From Psychosis: Systematic Review and Meta-Analysis 24 February 2021 | Frontiers in Psychiatry, Vol. 12
  • Mental Health 10 May 2021
  • Reciprocity membership: A potential pathway towards recovery from mental illness in a Middle Eastern context 5 December 2019 | Transcultural Psychiatry, Vol. 58, No. 1
  • Training mental health peer support training facilitators: a qualitative, participatory evaluation 6 September 2020 | International Journal of Mental Health Nursing, Vol. 30, No. 1
  • Benefits of Occupational Therapy Activity-Based Groups in a Singapore Acute Psychiatric Ward: Participants’ Perspectives 3 February 2021 | Occupational Therapy in Mental Health, Vol. 37, No. 1
  • Psychiatrists’ Perceptions of Schizophrenia and Its Recovery: A Thematic Analysis 1 July 2021
  • Internalized Stigma among Bedouin and Jews with Mental Illness: Comparing Self-Esteem, Hope, and Quality of Life 16 May 2020 | Psychiatric Quarterly, Vol. 91, No. 4
  • Narrative Matters: Mental health recovery – considerations when working with youth 1 September 2020 | Child and Adolescent Mental Health, Vol. 25, No. 4
  • Shared decision making for adults with severe mental illness: A concept analysis 5 August 2020 | Japan Journal of Nursing Science, Vol. 17, No. 4
  • Deprescribing in psychiatry: challenges and opportunities 25 September 2020 | Medical Herald of the South of Russia, Vol. 11, No. 3
  • The recovery model in chronic mental health: A community-based investigation of social identity processes Psychiatry Research, Vol. 291
  • Demographic and clinical profile of patients utilising a transitional care intervention in the Western Cape, South Africa 26 August 2020 | South African Journal of Psychiatry, Vol. 26
  • A framework to conceptualize personal recovery from eating disorders: A systematic review and qualitative meta‐synthesis of perspectives from individuals with lived experience 17 March 2020 | International Journal of Eating Disorders, Vol. 53, No. 8
  • Pathways to Recovery among Homeless People with Mental Illness: Is Impulsiveness Getting in the Way? 25 November 2019 | The Canadian Journal of Psychiatry, Vol. 65, No. 7
  • Recovery Programs for People With Mental Illness in Taiwan: A Feasibility Study 11 June 2020 | The American Journal of Occupational Therapy, Vol. 74, No. 4
  • “Entry to the Society from the Schizophrenic Cave”- A Qualitative Meta-Synthesis of Job Experiences for People with Schizophrenia 4 June 2020 | Issues in Mental Health Nursing, Vol. 5
  • A qualitative case study of a Korean former addict’s existential self‐interpretation and identity transformation 24 February 2020 | Asian Social Work and Policy Review, Vol. 14, No. 2
  • Leigh Evans , Ph.D. ,
  • Nancy J. Wewiorski , Ph.D. ,
  • Marsha Langer Ellison , Ph.D. ,
  • Pengsheng Ni , M.D., M.P.H. ,
  • Kimberly L. L. Harvey , M.P.H. ,
  • Marcia G. Hunt , Ph.D. ,
  • Jay A. Gorman , Ph.D. ,
  • Martin P. Charns , D.B.A.
  • Measuring Recovery in Deaf, Hard-of-Hearing, and Tinnitus Patients in a Mental Health Care Setting: Validation of the I.ROC 4 December 2019 | The Journal of Deaf Studies and Deaf Education, Vol. 25, No. 2
  • Ruhsal Hastalıklarda İyileşme: Kavram Analizi 31 March 2020 | Psikiyatride Guncel Yaklasimlar - Current Approaches in Psychiatry, Vol. 12, No. 1
  • An ambivalent atmosphere: Employment training programs and mental health recovery Health & Place, Vol. 62
  • Understanding the work of FLOs through a recovery framework lens 13 March 2018 | International Journal of Inclusive Education, Vol. 24, No. 1
  • Organization of Mental Health Services in Rural Areas 4 April 2020
  • Community and Mental Health 25 June 2020
  • Associations Between Two Domains of Social Adversity and Recovery Among Persons with Serious Mental Illnesses Being Treated in Community Mental Health Centers 24 September 2019 | Community Mental Health Journal, Vol. 56, No. 1
  • Social recovery: A neglected dimension of caring for women with perineal trauma in Iran Iranian Journal of Nursing and Midwifery Research, Vol. 25, No. 4
  • An Asian study on clinical and psychological factors associated with personal recovery in people with psychosis 22 August 2019 | BMC Psychiatry, Vol. 19, No. 1
  • The resource group method in severe mental illness: study protocol for a randomized controlled trial and a qualitative multiple case study 22 March 2019 | International Journal of Mental Health Systems, Vol. 13, No. 1
  • Antidepressant treatment with sertraline for adults with depressive symptoms in primary care: the PANDA research programme including RCT 1 December 2019 | Programme Grants for Applied Research, Vol. 7, No. 10
  • Similar yet unique: the victim's journey after acute sexual assault and the importance of continuity of care 9 May 2019 | Scandinavian Journal of Caring Sciences, Vol. 33, No. 4
  • The Forensic Restrictiveness Questionnaire: Development, Validation, and Revision 15 November 2019 | Frontiers in Psychiatry, Vol. 10
  • Luming Li , M.D. ,
  • Mengjie Deng , M.D., M.S. ,
  • Zhening Liu , M.D., Ph.D. ,
  • Robert Rohrbaugh , M.D.
  • ‘Monsters don’t bother me anymore’ forensic mental health service users’ experiences of acceptance and commitment therapy for psychosis 7 May 2019 | The Journal of Forensic Psychiatry & Psychology, Vol. 30, No. 4
  • Relasjonell recovery – utforsking avsamarbeid som bidrag til personers recovery i et botilbud Tidsskrift for psykisk helsearbeid, Vol. 16, No. 2
  • Psychiatric inpatient cost of care before and after admission at a residential subacute step-up/step-down mental health facility 15 March 2019 | Journal of Medical Economics, Vol. 22, No. 5
  • The Spring Foundation: a recovery approach to institutional public mental health services in South Africa 10 May 2019 | Perspectives in Public Health, Vol. 139, No. 3
  • Matthew Chinman , Ph.D. ,
  • Sharon McCarthy , Ph.D. ,
  • Chantele Mitchell-Miland , M.P.H. ,
  • Rachel L. Bachrach , Ph.D. ,
  • Russell K. Schutt , Ph.D. ,
  • Marsha Ellison , Ph.D.
  • Conceptualizations of subjective recovery from recent onset psychosis and its associated factors: A systematic review 21 June 2018 | Early Intervention in Psychiatry, Vol. 13, No. 2
  • Factor validation and Rasch analysis of the individual recovery outcomes counter 11 September 2017 | Disability and Rehabilitation, Vol. 41, No. 1
  • Re-orienting Mental Health Services to Mental Health Promotion 1 November 2019
  • Recovery in Mental Illness Among Rural Communities 30 May 2019
  • Journal de Thérapie Comportementale et Cognitive, Vol. 29, No. 2
  • Revista Colombiana de Psiquiatría, Vol. 48, No. 4
  • Revista Colombiana de Psiquiatría (English ed.), Vol. 48, No. 4
  • Journal of Mental Health, Vol. 28, No. 3
  • Social Work in Mental Health, Vol. 17, No. 3
  • Addiction Research & Theory, Vol. 27, No. 1
  • Critical Policy Studies, Vol. 13, No. 3
  • Qualitative Health Research, Vol. 29, No. 11
  • Promoting Recovery from Substance Misuse through Engagement with Community Assets: Asset Based Community Engagement 26 September 2019 | Substance Abuse: Research and Treatment, Vol. 13
  • Representations of mental health and arts participation in the national and local British press, 2007–2015 18 May 2017 | Health: An Interdisciplinary Journal for the Social Study of Health, Illness and Medicine, Vol. 23, No. 1
  • MODELS OF SOCIAL RECOVERY WITHIN THE CONTEXT OF PERSONAL-ORIENTED PARADIGM 1 January 2019 | Bulletin of Taras Shevchenko National University of Kyiv. Social work, No. 5
  • A new paradigm of youth recovery: Implications for youth mental health service provision 15 May 2018 | Australian Journal of Psychology, Vol. 70, No. 4
  • Mindfulness-based treatment of addiction: current state of the field and envisioning the next wave of research 18 April 2018 | Addiction Science & Clinical Practice, Vol. 13, No. 1
  • Introducing the recovery inspiration group: promoting hope for recovery with inspirational recovery stories Advances in Dual Diagnosis, Vol. 11, No. 4
  • Embedding recovery-based approaches into mental health nurse training British Journal of Mental Health Nursing, Vol. 7, No. 5
  • Risk profile of young people admitted to hospital for suicidal behaviour in Melbourne, Australia 20 May 2018 | Journal of Paediatrics and Child Health, Vol. 54, No. 11
  • Stability and mutual prospective relationships of stereotyped beliefs about mental illness, hope and depressive symptoms among people with schizophrenia spectrum disorders Psychiatry Research, Vol. 268
  • Why Frankenstein is a Stigma Among Scientists 26 June 2017 | Science and Engineering Ethics, Vol. 24, No. 4
  • Participatory arts, recovery and social inclusion Mental Health and Social Inclusion, Vol. 22, No. 3
  • Feasibility, Acceptability, and Preliminary Effectiveness of a Peer-Delivered and Technology Supported Self-Management Intervention for Older Adults with Serious Mental Illness 26 September 2017 | Psychiatric Quarterly, Vol. 89, No. 2
  • “Homeliness, hope and humour” (H 3 ) – ingredients for creating a therapeutic milieu in prisons Therapeutic Communities: The International Journal of Therapeutic Communities, Vol. 39, No. 1
  • Racial Differences in Mental Health Recovery among Veterans with Serious Mental Illness 14 April 2017 | Journal of Racial and Ethnic Health Disparities, Vol. 5, No. 2
  • Treatment and Spiritual Care in Mental Health Journal of Christian Nursing, Vol. 35, No. 2
  • Exploring the processes of change facilitated by musical activities on mental wellness 14 August 2017 | Nordic Journal of Music Therapy, Vol. 27, No. 2
  • Recovery concept in a Norwegian setting to be examined by the assertive community treatment model and mixed methods 27 December 2016 | International Journal of Mental Health Nursing, Vol. 27, No. 1
  • Administration and Policy in Mental Health and Mental Health Services Research, Vol. 45, No. 1
  • Community Mental Health Journal, Vol. 54, No. 8
  • Psychiatric Quarterly, Vol. 89, No. 2
  • Journal of Psychosocial Rehabilitation and Mental Health, Vol. 5, No. 1
  • Psychiatry Research, Vol. 267
  • Journal of Social Service Research, Vol. 44, No. 3
  • Mental Health, Religion & Culture, Vol. 21, No. 7
  • Advances in Mental Health, Vol. 16, No. 1
  • What Does Success Look Like in the Forensic Mental Health System? Perspectives of Service Users and Service Providers 21 March 2016 | International Journal of Offender Therapy and Comparative Criminology, Vol. 62, No. 1
  • BJPsych Advances, Vol. 24, No. 5
  • RIVISTA SPERIMENTALE DI FRENIATRIA, No. 3
  • International Journal of Environmental Research and Public Health, Vol. 15, No. 3
  • International Journal of Environmental Research and Public Health, Vol. 15, No. 7
  • Enhancing Spiritually Based Care Through Gratitude Practices: A Health-Care Improvement Project 1 January 2018 | Creative Nursing, Vol. 24, No. 1
  • Assessment of Needs: Differences between male and female patients with schizophrenia needs in psychiatric hospitals in Baghdad city 24 August 2017 | International Journal of Social Psychiatry, Vol. 63, No. 7
  • Recovery From Schizophrenia in Community-Based Psychosocial Rehabilitation Settings 15 June 2015 | Research on Social Work Practice, Vol. 27, No. 5
  • Enhancing Key Competencies of Health Professionals in the Assessment and Care of Adults at Risk of Suicide Through Education and Technology Clinical Nurse Specialist, Vol. 31, No. 5
  • The Sociology of Mental Health
  • Exploring personal recovery in mental illness through an Arabic sociocultural lens 18 November 2016 | Journal of Psychiatric and Mental Health Nursing, Vol. 24, No. 2-3
  • Outcomes of a care coordinated service model for persons with severe and persistent mental illness: A qualitative study 13 December 2016 | International Journal of Social Psychiatry, Vol. 63, No. 1
  • Preventive Self-Help and the Six Nonnaturals: Remedies from Burton’s Anatomy of Melancholy 28 February 2017
  • Community Mental Health Journal, Vol. 53, No. 3
  • Psychiatric Quarterly, Vol. 88, No. 3
  • Journal of Psychosocial Rehabilitation and Mental Health, Vol. 4, No. 2
  • Annales Médico-psychologiques, revue psychiatrique, Vol. 175, No. 7
  • General Hospital Psychiatry, Vol. 48
  • The Australian Journal of Rehabilitation Counselling, Vol. 23, No. 1
  • Alcoholism Treatment Quarterly, Vol. 35, No. 3
  • Scandinavian Journal of Occupational Therapy, Vol. 24, No. 3
  • International Journal of Mental Health Promotion, Vol. 19, No. 1
  • Journal of Dual Diagnosis, Vol. 13, No. 4
  • BMC Psychiatry, Vol. 17, No. 1
  • International Journal of Mental Health Systems, Vol. 11, No. 1
  • Scandinavian Psychologist, Vol. 4
  • Jornal Brasileiro de Psiquiatria, Vol. 66, No. 1
  • JMIR Medical Informatics, Vol. 5, No. 1
  • Open Journal of Psychiatry, Vol. 07, No. 01
  • Le Modèle global de santé mentale publique et les mentors de rétablissement 14 June 2017 | Santé mentale au Québec, Vol. 42, No. 1
  • Validity of clinically significant change classifications yielded by Jacobson-Truax and Hageman-Arrindell methods 6 June 2016 | BMC Psychiatry, Vol. 16, No. 1
  • Recovery from schizophrenia: developing context utilising the literature Mental Health and Social Inclusion, Vol. 20, No. 3
  • Recreation for mental health recovery 30 November 2016 | Leisure/Loisir, Vol. 40, No. 3
  • Why Do Adults With ADHD Choose Strength-Based Coaching Over Public Mental Health Care? A Qualitative Case Study From the Netherlands 8 August 2016 | SAGE Open, Vol. 6, No. 3
  • Treatment of an Older Adult With Borderline Personality Disorder and Prescription Opioid Abuse 8 January 2016 | Clinical Case Studies, Vol. 15, No. 3
  • Rural and remote communities, technology and mental health recovery Journal of Information, Communication and Ethics in Society, Vol. 14, No. 2
  • “Finding a way out”: Case histories of mental health care-seeking and recovery among long-term internally displaced persons in Georgia 23 December 2015 | Transcultural Psychiatry, Vol. 53, No. 2
  • Enhancing recovery orientation within mental health services: expanding the utility of values The Journal of Mental Health Training, Education and Practice, Vol. 11, No. 1
  • Understanding recovery: the perspective of substance misusing offenders Drugs and Alcohol Today, Vol. 16, No. 1
  • Family Interventions in Psychosis: A Review of the Evidence and Barriers to Implementation 12 November 2020 | Australian Psychologist, Vol. 51, No. 1
  • Evette J. Ludman , Ph.D. ,
  • Gregory E. Simon , M.D., M.P.H. ,
  • Louis C. Grothaus , M.A. ,
  • Julie Elissa Richards ,
  • Ursula Whiteside , Ph.D. ,
  • Christine Stewart , Ph.D.
  • Personality and Mental Health, Vol. 10, No. 4
  • The Recovery Model and Modern Psychiatric Care: Conceptual Perspective, Critical Approach and Practical Application 15 May 2016
  • International Journal of Mental Health and Addiction, Vol. 14, No. 5
  • Journal of Orthopaedics, Vol. 13, No. 2
  • Psychiatry Research, Vol. 245
  • International Journal of Mental Health, Vol. 45, No. 1
  • Australian Social Work, Vol. 69, No. 2
  • Child Care in Practice, Vol. 22, No. 1
  • BMC Psychiatry, Vol. 16, No. 1
  • International Journal of Mental Health Systems, Vol. 10, No. 1
  • PLOS ONE, Vol. 11, No. 8
  • Korean Journal of Health Psychology, Vol. 21, No. 1
  • Journal of Mental Health, Vol. 25, No. 4
  • Programme Grants for Applied Research, Vol. 4, No. 5
  • Frontiers in Psychology, Vol. 7
  • No More “Us” and “Them”: Integrating Recovery and Well-Being into a Conceptual Model for Mental Health Policy Canadian Journal of Community Mental Health, Vol. 34, No. 4
  • Development of Measures to Assess Personal Recovery in Young People Treated in Specialist Mental Health Services 10 June 2014 | Clinical Psychology & Psychotherapy, Vol. 22, No. 6
  • Advance directives in mental health care: evidence, challenges and promise 25 September 2015 | World Psychiatry, Vol. 14, No. 3
  • Recovery from psychosis: physical health, antipsychotic medication and the daily dilemmas for mental health nurses 2 August 2015 | Journal of Psychiatric and Mental Health Nursing, Vol. 22, No. 7
  • Theoretical Paper From shared roots to fruitful collaboration: How counselling psychology can benefit from (re)connecting with positive psychology 1 September 2015 | Counselling Psychology Review, Vol. 30, No. 3
  • Poles apart: does the export of mental health expertise from the Global North to the Global South represent a neutral relocation of knowledge and practice? 13 February 2015 | Sociology of Health & Illness, Vol. 37, No. 5
  • Evaluation of a recovery-oriented care training program for mental healthcare professionals: Effects on mental health consumer outcomes 25 June 2014 | International Journal of Social Psychiatry, Vol. 61, No. 2
  • Consumer Empowerment and the Essential Care Environment Ingredients 14 May 2015 | Journal of the American Psychiatric Nurses Association, Vol. 21, No. 2
  • Administration and Policy in Mental Health and Mental Health Services Research, Vol. 42, No. 2
  • Community Mental Health Journal, Vol. 51, No. 4
  • Community Mental Health Journal, Vol. 51, No. 1
  • Psychiatric Quarterly, Vol. 86, No. 4
  • International Journal of Mental Health and Addiction, Vol. 13, No. 5
  • Journal of Psychosocial Rehabilitation and Mental Health, Vol. 2, No. 1
  • Aggression and Violent Behavior, Vol. 22
  • Postępy Psychiatrii i Neurologii, Vol. 24, No. 2
  • Schizophrenia Research, Vol. 168, No. 1-2
  • Art Therapy, Vol. 32, No. 1
  • Nordic Journal of Music Therapy, Vol. 24, No. 3
  • Research in Dance Education, Vol. 16, No. 2
  • Social Work in Mental Health, Vol. 13, No. 5
  • Journal of Mental Health Research in Intellectual Disabilities, Vol. 8, No. 3-4
  • Journal of the Society for Social Work and Research, Vol. 6, No. 3
  • British Journal of Social Work, Vol. 45, No. suppl 1
  • Health and Social Welfare Review, Vol. 35, No. 3
  • Issues in Mental Health Nursing, Vol. 36, No. 8
  • Mental Health Management in Vocational Rehabilitation and Disability Evaluation: Applying the International Classification of Functioning, Disability and Health (ICF) Conceptual Framework 3 November 2014
  • Psychosoziale Therapien 24 May 2016
  • The recovery model and anorexia nervosa 13 June 2014 | Australian & New Zealand Journal of Psychiatry, Vol. 48, No. 11
  • Untangling the Nonrecyclable Citizen 21 August 2014 | Qualitative Health Research, Vol. 24, No. 10
  • Module 1: What is mental health recovery and how does it relate to person‐centered care planning? 20 June 2014
  • Adolescents with Anxiety and Depression: Is Social Recovery Relevant? 1 May 2013 | Clinical Psychology & Psychotherapy, Vol. 21, No. 4
  • ‘It’s like having a day of freedom, a day off from being ill’ : Exploring the experiences of people living with mental health problems who attend a community-based arts project, using interpretative phenomenological analysis 21 March 2013 | Journal of Health Psychology, Vol. 19, No. 6
  • A Research Tool for Mental Health Services Journal of Psychosocial Nursing and Mental Health Services, Vol. 52, No. 6
  • The theory–practice gap: Well and truly alive in mental health nursing 26 June 2014 | Nursing & Health Sciences, Vol. 16, No. 2
  • Hope and recovery: a scoping review Advances in Dual Diagnosis, Vol. 7, No. 2
  • Recovery and Severe Mental Illness: Description and Analysis 1 May 2014 | The Canadian Journal of Psychiatry, Vol. 59, No. 5
  • The significance of possible selves in patients of an early intervention programme for psychotic disorders 11 July 2013 | Early Intervention in Psychiatry, Vol. 8, No. 2
  • Psychometric evaluation of the Dutch version of the Mental Health Recovery Measure (MHRM) 7 February 2013 | International Journal of Social Psychiatry, Vol. 60, No. 2
  • Can we risk recovery? A grounded theory of clinical psychologists' perceptions of risk and recovery‐oriented mental health services 26 September 2012 | Psychology and Psychotherapy: Theory, Research and Practice, Vol. 87, No. 1
  • Administration and Policy in Mental Health and Mental Health Services Research, Vol. 41, No. 4
  • Community Mental Health Journal, Vol. 50, No. 4
  • Psychiatric Quarterly, Vol. 85, No. 2
  • Quality of Life Research, Vol. 23, No. 9
  • Journal of Psychosocial Rehabilitation and Mental Health, Vol. 1, No. 2
  • Social Science & Medicine, Vol. 103
  • Child & Youth Services, Vol. 35, No. 2
  • Australian Social Work, Vol. 67, No. 2
  • Alcoholism Treatment Quarterly, Vol. 32, No. 1
  • Journal of Human Behavior in the Social Environment, Vol. 24, No. 4
  • American Journal of Psychiatric Rehabilitation, Vol. 17, No. 1
  • Journal of the Society for Social Work and Research, Vol. 5, No. 2
  • British Journal of Social Work, Vol. 44, No. 5
  • Health Promotion International
  • Journal of Medicine and Philosophy, Vol. 39, No. 4
  • Clinical Psychology: Science and Practice, Vol. 21, No. 2
  • Shifting practices of recovery under community mental health reform: A street-level organizational ethnography 31 October 2013 | Qualitative Social Work, Vol. 13, No. 1
  • Journal of Child & Adolescent Mental Health, Vol. 26, No. 1
  • Journal of Mental Health, Vol. 23, No. 1
  • Intervention précoce pour la psychose : concepts, connaissances actuelles et orientations futures 15 December 2014 | Santé mentale au Québec, Vol. 39, No. 2
  • Everything Old Is New Again: Recovery and Serious Mental Illness 26 August 2014
  • Is risk assessment the new clinical model in public mental health? 29 July 2013 | Australasian Psychiatry, Vol. 21, No. 6
  • R ecovery E ducation: a tool for psychiatric nurses 28 February 2013 | Journal of Psychiatric and Mental Health Nursing, Vol. 20, No. 10
  • The Global Model of Public Mental Health through the WHO QualityRights project Journal of Public Mental Health, Vol. 12, No. 4
  • Gardening as a mental health intervention: a review Mental Health Review Journal, Vol. 18, No. 4
  • Professionals’ perceptions of the Mental Health Recovery Star Mental Health Review Journal, Vol. 18, No. 4
  • EMPOWERING VETERANS WITH PTSD IN THE RECOVERY ERA: ADVANCING DIALOGUE AND INTEGRATING SERVICES 2 September 2014 | Annals of Anthropological Practice, Vol. 37, No. 2
  • Sur la contribution des personnes utilisatrices de services de santé mentale en tant que partenaires d’enseignement en psychiatrie 28 August 2013 | Global Health Promotion, Vol. 20, No. 3
  • Facilitators and Barriers to Living with Psychosis: An Exploratory Collaborative Study of the Perspectives of Mental Health Service Users 1 September 2013 | British Journal of Occupational Therapy, Vol. 76, No. 9
  • Leopoldo J. Cabassa , Ph.D. ,
  • Andel Nicasio , M.S.Ed. , and
  • Rob Whitley , Ph.D.
  • Chinese translation and validation of the questionnaire on the process of recovery in schizophrenia and other psychotic disorders 6 June 2013 | Research in Nursing & Health, Vol. 36, No. 4
  • An Examination of the Feasibility of Adventure-Based Therapy in Outpatient Care for Individuals With Psychosis Canadian Journal of Community Mental Health, Vol. 32, No. 2
  • A recovery perspective on community day leaves The Journal of Forensic Practice, Vol. 15, No. 2
  • Developing an Evaluation Framework for Consumer-Centred Collaborative Care of Depression Using Input from Stakeholders 1 March 2013 | The Canadian Journal of Psychiatry, Vol. 58, No. 3
  • Administration and Policy in Mental Health and Mental Health Services Research, Vol. 40, No. 3
  • Community Mental Health Journal, Vol. 49, No. 3
  • Psychiatry Research, Vol. 209, No. 3
  • Psychiatry Research, Vol. 208, No. 1
  • Social Science & Medicine, Vol. 99
  • International Journal of Forensic Mental Health, Vol. 12, No. 2
  • Psychoanalytic Social Work, Vol. 20, No. 1
  • Social Work in Mental Health, Vol. 11, No. 4
  • American Journal of Psychiatric Rehabilitation, Vol. 16, No. 2
  • Social Work Research, Vol. 37, No. 3
  • Southern Medical Journal, Vol. 106, No. 1
  • The Journal of Digital Policy and Management, Vol. 11, No. 12
  • Journal of Systemic Therapies, Vol. 32, No. 2
  • Journal of Systemic Therapies, Vol. 32, No. 4
  • Spirit Lifting: Hope and Recovery in Case Management Practice 3 May 2018 | Families in Society: The Journal of Contemporary Social Services, Vol. 94, No. 1
  • International Journal of Self Help and Self Care, Vol. 7, No. 1
  • Health, Vol. 05, No. 03
  • Community Psychiatry: What Should Future Psychiatrists Learn? Harvard Review of Psychiatry, Vol. 20, No. 6
  • What Do Psychotic Experiences Mean to Chinese Schizophrenia Patients? 3 October 2012 | Qualitative Health Research, Vol. 22, No. 12
  • Teaching Supportive Psychotherapy in the Twenty-First Century Harvard Review of Psychiatry, Vol. 20, No. 5
  • Issues and developments on the consumer recovery construct 12 March 2013 | World Psychiatry, Vol. 11, No. 3
  • Profiles of individually defined recovery of people with major psychiatric problems 3 August 2011 | International Journal of Social Psychiatry, Vol. 58, No. 5
  • The recovery model and complex health needs: What health psychology can learn from mental health and substance misuse service provision 21 October 2011 | Journal of Health Psychology, Vol. 17, No. 5
  • Deborah Windell , M.Sc. ,
  • Ross Norman , Ph.D. , and
  • Ashok K. Malla , M.B.B.S., F.R.C.P.C.
  • Community mental health nurses' perspectives of recovery‐oriented practice 27 October 2011 | Journal of Psychiatric and Mental Health Nursing, Vol. 19, No. 4
  • Minimizing and treating chronicity in the eating disorders: A clinical overview 23 January 2012 | International Journal of Eating Disorders, Vol. 45, No. 4
  • Restorative Integral Support (RIS) for Older Adults Experiencing Co-Occurring Disorders 17 May 2012 | The International Journal of Aging and Human Development, Vol. 74, No. 3
  • Hanneke van Gestel-Timmermans , Ph.D. ,
  • Evelien P. M. Brouwers , Ph.D. ,
  • Marcel A. L. M. van Assen , Ph.D. , and
  • Chijs van Nieuwenhuizen , Ph.D.
  • Social Psychiatry and Psychiatric Epidemiology, Vol. 47, No. 7
  • Administration and Policy in Mental Health and Mental Health Services Research, Vol. 39, No. 3
  • Current Psychiatry Reports, Vol. 14, No. 3
  • Archives of Psychiatric Nursing, Vol. 26, No. 4
  • International Journal of Nursing Studies, Vol. 49, No. 11
  • Social Science & Medicine, Vol. 74, No. 2
  • Irish Journal of Psychological Medicine, Vol. 29, No. 3
  • Social Work in Health Care, Vol. 51, No. 1
  • Journal of Prevention & Intervention in the Community, Vol. 40, No. 4
  • Person-Centered & Experiential Psychotherapies, Vol. 11, No. 4
  • American Journal of Psychiatric Rehabilitation, Vol. 15, No. 3
  • Arts & Health, Vol. 4, No. 1
  • The Scientific World Journal, Vol. 2012
  • Nursing Research and Practice, Vol. 2012
  • International Journal of Mental Health Systems, Vol. 6, No. 1
  • Journal of Korean Academy of Psychiatric and Mental Health Nursing, Vol. 21, No. 1
  • Journal of Korean Academy of Psychiatric and Mental Health Nursing, Vol. 21, No. 3
  • Saúde e Sociedade, Vol. 21, No. 3
  • International Journal of Self Help and Self Care, Vol. 6, No. 2
  • Psychiatric Rehabilitation Journal, Vol. 35, No. 4
  • Anthropological Perspectives on Money Management: Considerations for the Design and Implementation of Interventions for Substance Abuse 19 December 2011 | The American Journal of Drug and Alcohol Abuse, Vol. 38, No. 1
  • Issues in Mental Health Nursing, Vol. 33, No. 6
  • International Review of Psychiatry, Vol. 24, No. 1
  • Journal of Mental Health, Vol. 21, No. 2
  • Clinical Schizophrenia & Related Psychoses, Vol. 5, No. 4
  • Se rétablir de troubles psychiatriques : un changement de regard sur le devenir des personnes L'information psychiatrique, Vol. 88, No. 4
  • Iryo To Shakai, Vol. 22, No. 2
  • Clair Le Boutillier , B.Sc., M.Sc. ,
  • Mary Leamy , M.Sc., Ph.D. ,
  • Victoria J. Bird , B.Sc. ,
  • Larry Davidson , Ph.D. ,
  • Julie Williams , B.Sc., M.Sc. , and
  • Mike Slade , Psych.D., Ph.D.
  • Research International Journal of Therapy and Rehabilitation, Vol. 18, No. 12
  • Commentaries International Journal of Therapy and Rehabilitation, Vol. 18, No. 12
  • The recovery paradigm in forensic mental health services 23 November 2011 | Criminal Behaviour and Mental Health, Vol. 21, No. 5
  • A substantive theory of recovery from the effects of severe persistent mental illness 21 July 2010 | International Journal of Social Psychiatry, Vol. 57, No. 6
  • There seems no place for place: a gap analysis of the recovery literature Journal of Public Mental Health, Vol. 10, No. 3
  • A UK validation of the Stages of Recovery Instrument 8 April 2010 | International Journal of Social Psychiatry, Vol. 57, No. 5
  • Stigma, Reflected Appraisals, and Recovery Outcomes in Mental Illness 10 June 2011 | Social Psychology Quarterly, Vol. 74, No. 2
  • Quantitative relationship between recovery and benefit‐finding among persons with chronic mental illness in Japan 12 April 2011 | Nursing & Health Sciences, Vol. 13, No. 2
  • A service evaluation of recovery support from the patient’s perspective British Journal of Wellbeing, Vol. 2, No. 5
  • Home Bittersweet Home: the Significance of Home for Occupational Transformations 27 July 2016 | International Journal of Social Psychiatry, Vol. 57, No. 3
  • Creating Occupational Engagement to Maximise Recovery in Mental Health 11 October 2013
  • Quality of Life Research, Vol. 20, No. 7
  • Revista de Psiquiatría y Salud Mental (English Edition), Vol. 4, No. 1
  • Archives of Psychiatric Nursing, Vol. 25, No. 5
  • International Journal of Orthopaedic and Trauma Nursing, Vol. 15, No. 1
  • Psychiatry Research, Vol. 189, No. 3
  • Revista de Psiquiatría y Salud Mental, Vol. 4, No. 1
  • Irish Journal of Psychological Medicine, Vol. 28, No. 3
  • Journal of Advanced Nursing, Vol. 67, No. 10
  • The lived experience of art making as a companion to the mental health recovery process 9 August 2010 | Disability and Rehabilitation, Vol. 33, No. 8
  • Rob Whitley, Ph.D.
  • Robert E. Drake, M.D., Ph.D.
  • Not Just Talk: What a Psychosocial Primary Care Mental Health Service Can Do Practice, Vol. 22, No. 5
  • Hope as a determinant of mental health recovery: a psychometric evaluation of the Herth Hope Index‐Dutch version 11 November 2010 | Scandinavian Journal of Caring Sciences, Vol. 24, No. s1
  • Approaches to Developing OPB Formulations 14 March 2011
  • Reliability and validity of the Japanese version of the Self‐Identified Stage of Recovery for people with long term mental illness 5 May 2010 | International Journal of Mental Health Nursing, Vol. 19, No. 3
  • Principles of Social Intervention 28 March 2010
  • Social Interventions for Psychosis 28 March 2010
  • When Symptoms and Treatments Hinder Vocational Recovery 16 February 2010
  • Reliability and validity of the Japanese version of the Recovery Assessment Scale (RAS) for people with chronic mental illness: Scale development International Journal of Nursing Studies, Vol. 47, No. 3
  • Exploring quality of life in the eating disorders 2 December 2009 | European Eating Disorders Review, Vol. 18, No. 2
  • Confusion of recovery: One solution International Journal of Mental Health Nursing, Vol. 19, No. 1
  • Shared decision making interventions for people with mental health conditions 20 January 2010 | Cochrane Database of Systematic Reviews, Vol. 68
  • Michele R. Spoont, Ph.D.
  • Maureen Murdoch, M.D., M.P.H.
  • James Hodges, Ph.D., M.A.
  • Sean Nugent, B.A.
  • Culture, Medicine, and Psychiatry, Vol. 34, No. 3
  • Theoretical Medicine and Bioethics, Vol. 31, No. 1
  • International Journal of Mental Health and Addiction, Vol. 8, No. 2
  • Schizophrenia Research, Vol. 119, No. 1-3
  • Schizophrenia Bulletin, Vol. 36, No. 1
  • American Journal of Orthopsychiatry, Vol. 80, No. 2
  • American Journal of Orthopsychiatry, Vol. 80, No. 3
  • Coming Back Normal: The Process of Self-Recovery in Those With Schizophrenia 21 January 2010 | Journal of the American Psychiatric Nurses Association, Vol. 16, No. 1
  • BMC Health Services Research, Vol. 10, No. 1
  • Journal of Korean Academy of Psychiatric and Mental Health Nursing, Vol. 19, No. 2
  • Journal of Korean Academy of Psychiatric and Mental Health Nursing, Vol. 19, No. 3
  • Psychiatry: Interpersonal and Biological Processes, Vol. 73, No. 1
  • International Review of Psychiatry, Vol. 22, No. 2
  • Journal of Mental Health, Vol. 19, No. 1
  • Disability and Rehabilitation, Vol. 32, No. 12
  • The Recovery Process Utilizing Erikson’s Stages of Human Development 17 June 2009 | Community Mental Health Journal, Vol. 45, No. 6
  • Treatment effect and recovery — dilemmas in dual diagnosis treatment 12 April 2017 | Nordic Studies on Alcohol and Drugs, Vol. 26, No. 6
  • Relationships Between Humor, Subversion, and Genuine Connection Among Persons With Severe Mental Illness 5 October 2009 | Qualitative Health Research, Vol. 19, No. 10
  • Recovery in the Canadian Context: Feedback on the Framework for Mental Health Strategy Development Canadian Journal of Community Mental Health, Vol. 28, No. 2
  • The Creation of “We Are Neighbours”: Participatory Research and Recovery Canadian Journal of Community Mental Health, Vol. 28, No. 2
  • Beyond Mental Health Maintenance: An Evaluation Framework Driven by Recovery-Focused Outcomes Canadian Journal of Community Mental Health, Vol. 28, No. 2
  • Skye Barbic, M.Sc.
  • Terry Krupa, Ph.D.
  • Irene Armstrong, Ph.D.
  • Experiential Knowledge of Serious Mental Health Problems 3 June 2008 | Journal of Humanistic Psychology, Vol. 49, No. 2
  • Robert Hunter, M.D., F.R.C.Psych.
  • Rosie Cameron, B.A., M.B.A.
  • John Norrie, B.Sc., M.Sc.
  • Community Mental Health Journal, Vol. 45, No. 3
  • Psychiatry Research, Vol. 167, No. 3
  • Irish Journal of Psychological Medicine, Vol. 26, No. 4
  • Issues in Mental Health Nursing, Vol. 30, No. 8
  • Occupational Therapy in Mental Health, Vol. 25, No. 2
  • Journal of Social Work Practice, Vol. 23, No. 4
  • Journal of Mental Health, Vol. 18, No. 2
  • Substance Use & Misuse, Vol. 44, No. 4
  • International Journal of Mental Health Promotion, Vol. 11, No. 4
  • Social Work in Mental Health, Vol. 7, No. 6
  • Psychosis, Vol. 1, No. 2
  • Mental Health Review Journal, Vol. 14, No. 4
  • BMC Health Services Research, Vol. 9, No. 1
  • The Australian Journal of Rehabilitation Counselling, Vol. 15, No. 1
  • Korean Journal of Clinical Psychology, Vol. 28, No. 2
  • Journal of Japan Academy of Nursing Science, Vol. 29, No. 3
  • `Recovery' and current mental health policy 1 December 2008 | Chronic Illness, Vol. 4, No. 4
  • Creating a space for recovery-focused psychiatric nursing care Nursing Inquiry, Vol. 15, No. 3
  • Depression and other mental illnesses in the context of workplace efficacy 29 July 2008 | Journal of Psychiatric and Mental Health Nursing, Vol. 15, No. 7
  • Approaching in the right spirit: Spirituality and hope in recovery from mental health problems International Journal of Therapy and Rehabilitation, Vol. 15, No. 6
  • Clinical problems in the long-term care of patients with chronic depression Journal of Advanced Nursing, Vol. 62, No. 6
  • Jeffrey A. Lieberman, M.D.
  • Lloyd I. Sederer, M.D.
  • Aysenil Belger, Ph.D.
  • Richard Keefe, Ph.D.
  • Diana Perkins, M.D., M.P.H.
  • Scott Stroup, M.D., M.P.H.
  • Professional Differences in Attitudes toward and Utilization of Psychiatric Recovery 9 May 2018 | Families in Society: The Journal of Contemporary Social Services, Vol. 89, No. 2
  • Determining the effectiveness of mental health services from a consumer perspective: Part 2: Barriers to recovery and principles for evaluation 28 February 2008 | International Journal of Mental Health Nursing, Vol. 17, No. 2
  • Determining the effectiveness of mental health services from a consumer perspective: Part 1: Enhancing recovery 28 February 2008 | International Journal of Mental Health Nursing, Vol. 17, No. 2
  • Recovery-Based Practice: Do We Know What We Mean or Mean What We Know? 1 January 2008 | Australian & New Zealand Journal of Psychiatry, Vol. 42, No. 3
  • Using Qualitative Research to Inform Mental Health Policy 1 March 2008 | The Canadian Journal of Psychiatry, Vol. 53, No. 3
  • Existential struggle and self-reported needs of patients in rehabilitation 24 January 2008 | Journal of Advanced Nursing, Vol. 61, No. 4
  • Jules M. Ranz, M.D.
  • Anthony D. Mancini, Ph.D.
  • Norma C. Ware, Ph.D.
  • Kim Hopper, Ph.D.
  • Toni Tugenberg, M.Ed., L.I.C.S.W.
  • Barbara Dickey, Ph.D.
  • Daniel Fisher, M.D., Ph.D.
  • The Journal of Behavioral Health Services & Research, Vol. 35, No. 4
  • The Doctor-Patient Relationship
  • Clinical Psychology Review, Vol. 28, No. 7
  • International Journal of Nursing Studies, Vol. 45, No. 8
  • Social Theory & Health, Vol. 6, No. 4
  • Journal of Teaching in the Addictions, Vol. 7, No. 2
  • Journal of Dual Diagnosis, Vol. 4, No. 3
  • Psychopathology, Vol. 41, No. 5
  • Advances in Psychiatric Treatment, Vol. 14, No. 5
  • Advances in Psychiatric Treatment, Vol. 14, No. 4
  • Groupwork, Vol. 18, No. 1
  • Catching life: the contribution of arts initiatives to recovery approaches in mental health 22 November 2007 | Journal of Psychiatric and Mental Health Nursing, Vol. 14, No. 8
  • The recovery alliance theory of mental health nursing 22 November 2007 | Journal of Psychiatric and Mental Health Nursing, Vol. 14, No. 8
  • Miriam C. Tepper, M.D.
  • Joseph A. Rogers, A.A.
  • Michael J. Vergare, M.D.
  • Richard C. Baron, M.A.
  • Mark S. Salzer, Ph.D.
  • Guide Schizophrenia Patients to Better Physical Health The Nurse Practitioner, Vol. 32, No. 7
  • Role of psychological factors in studying recovery from a transactional stress‐coping approach: Implications for mental health nursing practices 22 May 2007 | International Journal of Mental Health Nursing, Vol. 16, No. 3
  • Art and Recovery in Mental Health: A Qualitative Investigation 5 November 2016 | British Journal of Occupational Therapy, Vol. 70, No. 5
  • The Recovery Process and People with Serious Mental Illness Living in the Community 8 September 2008 | Occupational Therapy in Mental Health, Vol. 23, No. 2
  • The Concept of Recovery as an Organizing Principle for Integrating Mental Health and Addiction Services 10 March 2007 | The Journal of Behavioral Health Services & Research, Vol. 34, No. 2
  • Michael Ussher, Ph.D., M.Sc.
  • Liam Stanbury, M.B.B.S., B.Sc.
  • Vicky Cheeseman, M.B.B.S., B.Sc.
  • Guy Faulkner, Ph.D., M.Sc.
  • The Role of Self–efficacy in Recovery from Serious Psychiatric Disabilities: A Qualitative Study with Fifteen Psychiatric Survivors 15 August 2016 | Qualitative Social Work, Vol. 6, No. 1
  • Community Mental Health Journal, Vol. 43, No. 6
  • Archives of Psychiatric Nursing, Vol. 21, No. 6
  • Archives of Psychiatric Nursing, Vol. 21, No. 3
  • Social Science & Medicine, Vol. 64, No. 9
  • Epidemiologia e Psichiatria Sociale, Vol. 16, No. 3
  • Journal of Mental Health, Vol. 16, No. 4
  • The British Journal of Forensic Practice, Vol. 9, No. 4
  • Psychiatric Bulletin, Vol. 31, No. 2
  • Psychiatric Bulletin, Vol. 31, No. 9
  • Journal of Human Behavior in the Social Environment, Vol. 15, No. 2-3
  • Psychiatric Rehabilitation Journal, Vol. 31, No. 1
  • Clinical Schizophrenia & Related Psychoses, Vol. 1, No. 1
  • Biomédica, Vol. 27, No. 2
  • Developing a Comprehensive Understanding of the Working Alliance in Community Mental Health 1 July 2016 | Qualitative Health Research, Vol. 16, No. 8
  • A Solution-Focused Approach to Case Management and Recovery with Consumers who Have a Severe Mental Disability 3 May 2018 | Families in Society: The Journal of Contemporary Social Services, Vol. 87, No. 3
  • Challenges to Implementing and Sustaining Comprehensive Mental Health Service Programs 24 June 2016 | Evaluation & the Health Professions, Vol. 29, No. 2
  • Larry Davidson, Ph.D.
  • Maria O'Connell, Ph.D.
  • Janis Tondora, Psy.D.
  • Thomas Styron, Ph.D.
  • Karen Kangas, Ed.D.
  • Journal of Community Psychology, Vol. 34, No. 3
  • Administration and Policy in Mental Health and Mental Health Services Research, Vol. 33, No. 4
  • The Journal of Behavioral Health Services & Research, Vol. 33, No. 2
  • The Journal of Behavioral Health Services & Research, Vol. 33, No. 4
  • Health and Social Care in the Community, Vol. 14, No. 2
  • British Journal of Psychiatry, Vol. 188, No. 5
  • Journal of Teaching in the Addictions, Vol. 5, No. 2
  • Psychiatry: Interpersonal and Biological Processes, Vol. 69, No. 2
  • Federal Sentencing Reporter, Vol. 19, No. 2
  • Subjective Experience of Recovery from Schizophrenia-Related Disorders and Atypical Antipsychotics 27 July 2016 | International Journal of Social Psychiatry, Vol. 51, No. 3
  • Recovery in Community: Using Participatory Action Research to Explore Recovery with Alternatives Canadian Journal of Community Mental Health, Vol. 24, No. 2
  • General Practitioners and Mental Health Staff Sharing Patient Care: Working Model 25 June 2016 | Australasian Psychiatry, Vol. 13, No. 2
  • The Psychological Record, Vol. 55, No. 3
  • American Journal of Community Psychology, Vol. 36, No. 3-4
  • Community Mental Health Journal, Vol. 41, No. 1
  • Community Mental Health Journal, Vol. 41, No. 2
  • Community Mental Health Journal, Vol. 41, No. 3
  • Schizophrenia Research, Vol. 75, No. 1
  • Professional Psychology: Research and Practice, Vol. 36, No. 5
  • American Journal of Psychiatric Rehabilitation, Vol. 8, No. 3
  • Mental Health Review Journal, Vol. 10, No. 2
  • A Life in the Day, Vol. 9, No. 4
  • Psychiatry, Vol. 4, No. 11
  • Psychiatric Rehabilitation Journal, Vol. 28, No. 3
  • Psychiatric Rehabilitation Journal, Vol. 28, No. 4
  • Psychiatric Rehabilitation Journal, Vol. 29, No. 1
  • Entendre des voix : nouvelles voies ouvrant sur la pratique et la recherche 3 August 2005 | Santé mentale au Québec, Vol. 30, No. 1
  • What can Community Support Programs Do to Promote Productivity?: Perspectives of Service Users Canadian Journal of Community Mental Health, Vol. 23, No. 2
  • Sandra G. Resnick , Ph.D. ,
  • Robert A. Rosenheck , M.D. , and
  • Anthony F. Lehman , M.D.
  • Psychiatric Rehabilitation Journal, Vol. 28, No. 1
  • The Canadian Occupational Performance Measure: its use with clients with schizophrenia British Journal of Therapy and Rehabilitation, Vol. 10, No. 12
  • The Experience of Recovery from Schizophrenia: Towards an Empirically Validated Stage Model 17 November 2016 | Australian & New Zealand Journal of Psychiatry, Vol. 37, No. 5
  • Predictors of psychological distress in family caregivers of persons with psychiatric disabilities 4 September 2003 | Journal of Psychiatric and Mental Health Nursing, Vol. 10, No. 5
  • On the path to recovery: Patients’ experiences of treatment with long‐acting injections of antipsychotic medication 15 May 2003 | International Journal of Mental Health Nursing, Vol. 12, No. 2
  • Defining Recovery: An Interactionist Analysis of Mental Health Policy Development, Wisconsin 1996-1999 16 November 2016 | Qualitative Health Research, Vol. 13, No. 3
  • Social Service Review, Vol. 77, No. 2
  • Journal of Psychiatric Practice, Vol. 9, No. 2
  • Psychiatric Rehabilitation Journal, Vol. 27, No. 1
  • A Theory of Recovery Journal of Psychosocial Nursing and Mental Health Services, Vol. 40, No. 12
  • Frederick J. FreseIII , Ph.D. ,
  • Jonathan Stanley , J.D. ,
  • Ken Kress , J.D., Ph.D. , and
  • Suzanne Vogel-Scibilia , M.D.
  • Burton Hutto , M.D.
  • Herbert S. Peyser , M.D.

recovery model research paper

  • Bipolar Disorder
  • Therapy Center
  • When To See a Therapist
  • Types of Therapy
  • Best Online Therapy
  • Best Couples Therapy
  • Best Family Therapy
  • Managing Stress
  • Sleep and Dreaming
  • Understanding Emotions
  • Self-Improvement
  • Healthy Relationships
  • Student Resources
  • Personality Types
  • Guided Meditations
  • Verywell Mind Insights
  • 2023 Verywell Mind 25
  • Mental Health in the Classroom
  • Editorial Process
  • Meet Our Review Board
  • Crisis Support

The Recovery Model in Mental Health Care

Person-Centered, Holistic Approach

 Sarah Lyon, OTR/L, is a board-certified occupational therapist and founder of OT Potential.

recovery model research paper

  • Recovery Is Possible
  • Patient-Directed
  • Key Elements
  • Push for Recovery
  • Recovery Model vs. Medical Model
  • Limitations

Frequently Asked Questions

The recovery model is a holistic, person-centered approach to mental health care . The model has quickly gained momentum and is becoming the standard model of mental health care. It is based on two simple premises:

  • It is possible to recover from a mental health condition.
  • The most effective recovery is patient-directed.

If you’re receiving mental health services or have a loved one with a mental health condition, knowing the basic tenets of this model can help you advocate for the best care.

The framework can give you language to use when describing gaps in service. Your input can be invaluable in helping mental health care providers shift toward the values outlined by this model.

The Recovery Model Suggests Recovery Is Possible

The hallmark principle of the recovery model of mental health is the belief that people can recover from mental illness to lead full, satisfying lives. Until the mid-1970s, many practitioners believed that patients with mental health conditions were doomed to live with their illness forever and would not be able to contribute to society.

This belief particularly affected people with schizophrenia , schizoaffective disorder , and bipolar disorder . However, several long-term studies from several countries, published in the mid-70s, showed this to be false.

The recovery model is used in occupational therapy, a treatment type for both physical and mental health that focuses on the "client-provider partnership" and allows clients to choose what works best for their recovery.

You will also see elements of the recovery model in social work theory, where values such as client self-determination and well-being are emphasized.

What Are Recovery Goals?

The goals of the recovery model include helping people look beyond the limitations of their mental health conditions, encouraging them to strive for and achieve personal ambitions, and inspiring them to create meaningful relationships and personal connections.

The Recovery Model Is Patient-Directed

Often, sound evidence is not enough to change systems. It took two decades for this basic belief to gain traction in the medical community. The change came about largely through patients advocating to be involved in their own treatment .

Patients also began showing, through lived experience, that given the proper supports, they could live active lives in the community. The history of the movement reflects the second basic pillar of the recovery model: The most lasting change happens when the patient directs it.

Characteristics of the Recovery Model

The recovery model of mental health takes a holistic view of a person’s life. The Substance Abuse and Mental Health Services Administration (SAMHSA) defines recovery from mental disorders and/or substance use disorders as "a process of change through which individuals improve their health and wellness, live a self-directed life, and strive to reach their full potential."

SAMHSA outlines four dimensions that support recovery:

  • Health : In order to manage or recover from mental illness, people must make choices that support both their physical and mental well-being.
  • Home : People need a safe and stable place to live.
  • Purpose : Meaningful daily routines such as school, work, family, and community participation are important during the recovery process and for maintaining wellness.
  • Community : Supportive social relationships provide people with the love, emotional availability, and respect that they need to survive and thrive.

In particular, the recovery model of mental health stresses the importance of connectedness and social support . When people have supportive relationships that offer unconditional love, they are better able to cope with the symptoms of their illness and work toward recovery.

Psychologists , psychiatrists, doctors, and other health professionals can provide such support to a certain degree, but connections offered by friends, family, and other peers are also critical. Support groups and community organizations can help fulfill this need as well.

Principles of Treatment

SAMHSA also defines ten guiding principles for recovery treatment. Every institution that operates according to the recovery model should be striving to incorporate these into their care. The 10 core elements of the recovery model are:

  • Emerges from hope
  • Is person-driven
  • Occurs through many pathways
  • Is holistic
  • Is supported by peers and allies
  • Is supported through relationships and social networks
  • Is culturally based and influenced
  • Is supported by addressing trauma
  • Involves individual, family, and community strengths and responsibility
  • Is based on respect

The National Push for Recovery

By 2003, individuals who had been advocating for recovery-based care found their work paying off. A mental health commission appointed by President George W. Bush gave the final report of its work and made recovery-based care a national priority. This final report was ambitious. It envisioned a future that focused on the prevention, early detection, and cure of mental illness.

Today, the concept of the recovery model is familiar to most mental health practitioners. But individuals are still working out how to design programs and treatments based on these principles.

For an in-depth look at the recovery model, the American Psychological Association has 15 learning modules that are accessible to the public. The topics range from a broad overview of the recovery model to ways it is being implemented in practice. 

The Recovery Model vs. the Medical Model

The recovery model of mental illness is often contrasted against what is known as the medical model . The medical model posits that mental disorders have physiological causes, so the focus is often on the use of medications for treatment.

While the two models are often presented as being in opposition to one another, researchers have suggested that they are complementary and can be used together.

The medical model ensures that biological causes are fully addressed and that people receive the medication-based treatments that they need, while the recovery model ensures that patients are able to be directly involved in their own treatment.

The medical model is rooted in using treatments that are based on empirical research. The recovery model offers the personal empowerment and peer support that people need to cope with their illness and work toward getting better. A number of programs, including the Wellness Recovery Action Plan and the NAMI Family-to-Family program, incorporate both models and have research to back their effectiveness.

Limitations of the Recovery Model

While there are benefits to creating a unique healing program based on someone's subjective experience of their illness, there are potential drawbacks to using the recovery model.

Because the recovery model is not one consistent program (its components vary based on the client receiving treatment), it can be difficult to measure its outcomes or effectiveness.

In addition, some mental health conditions make it more difficult for a person to participate in guiding their own treatment plan. For instance, some people experiencing psychosis may not view themselves as having a mental illness.

In other cases, a person's symptoms might be so distressing that they require immediate medical attention. In this situation, the person experiencing mental illness cannot contribute to or make suggestions for their healthcare plan until their symptoms are addressed.

A Word From Verywell

One of the major strengths of the recovery model is that it focuses on individual strengths and abilities rather than on deficits and pathologies. It places trust in the individual to know their own experience and to be able to take an active role in their treatment.

In therapy, the recovery model emphasizes the important of self-determination, responsiblity, hope, and dignity. It sugguests that people can recovery from mental illness and that the goal of therapy is to help people achieve their fullest potential.

The recovery model of mental health focuses on empowering people to make decisions about their own lives and mental health. The core beliefs of this model are that it is possible to recover from mental illness and this recovery should be self-directed.

Jacob KS. Recovery model of mental illness: A complementary approach to psychiatric care . Indian J Psychol Med . 2015;37(2):117-119. doi:10.4103/0253-7176.155605

Malla A, Joober R, Garcia A. Mental illness is like any other medical illness: A critical examination of the statement and its impact on patient care and society . J Psychiatry Neurosci. 2015;40(3):147-150. doi:10.1503/jpn.150099

American Occupational Therapy Association. Occupational therapy's role with mental health recovery .

Webber M, Joubert L. Social work and recovery . Br J Soc Work. 2015;45(suppl 1):i1-i8. doi:10.1093/bjsw/bcv125

Substance Abuse and Mental Health Services Administration. Recovery and recovery support .

Hogan MF. The President's New Freedom Commission: Recommendations to transform mental health care in America . Psychiatr Serv . 2003;54(11):1467-1474. doi:10.1176/appi.ps.54.11.1467

National Alliance on Mental Illness. Science meets the human experience: Integrating the medical and recovery models .

By Sarah Lyon, OTR/L Sarah Lyon, OTR/L, is a board-certified occupational therapist and founder of OT Potential.

  • Open access
  • Published: 15 November 2022

Conceptualizing eating disorder recovery research: Current perspectives and future research directions

  • Heather Hower 1 , 2 ,
  • Andrea LaMarre 3 ,
  • Rachel Bachner-Melman 4 , 5 ,
  • Erin N. Harrop 6 ,
  • Beth McGilley 7 &
  • Therese E. Kenny 8  

Journal of Eating Disorders volume  10 , Article number:  165 ( 2022 ) Cite this article

4387 Accesses

3 Citations

6 Altmetric

Metrics details

How we research eating disorder (ED) recovery impacts what we know (perceive as fact) about it. Traditionally, research has focused more on the “what” of recovery (e.g., establishing criteria for recovery, reaching consensus definitions) than the “how” of recovery research (e.g., type of methodologies, triangulation of perspectives). In this paper we aim to provide an overview of the ED field’s current perspectives on recovery, discuss how our methodologies shape what is known about recovery, and suggest a broadening of our methodological “toolkits” in order to form a more complete picture of recovery.

This paper examines commonly used methodologies in research, and explores how incorporating different perspectives can add to our understanding of the recovery process. To do this, we (1) provide an overview of commonly used methodologies (quantitative, qualitative), (2) consider their benefits and limitations, (3) explore newer approaches, including mixed-methods, creative methods (e.g., Photovoice, digital storytelling), and multi-methods (e.g., quantitative, qualitative, creative methods, psycho/physiological, behavioral, laboratory, online observations), and (4) suggest that broadening our methodological “toolkits” could spur more nuanced and specific insights about ED recoveries. We propose a potential future research model that would ideally have a multi-methods design, incorporate different perspectives (e.g., expanding recruitment of diverse participants, including supportive others, in study co-creation), and a longitudinal course (e.g., capturing cognitive and emotional recovery, which often comes after physical). In this way, we hope to move the field towards different, more comprehensive, perspectives on ED recovery.

Our current perspectives on studying ED recovery leave critical gaps in our knowledge about the process. The traditional research methodologies impact our conceptualization of recovery definitions, and in turn limit our understanding of the phenomenon. We suggest that we expand our range of methodologies, perspectives, and timeframes in research, in order to form a more complete picture of what is possible in recovery; the multiple aspects of an individual’s life that can improve, the greater number of people who can recover than previously believed, and the reaffirmation of hope that, even after decades, individuals can begin, and successfully continue, their ED recovery process.

Plain English summary

How we research eating disorder (ED) recovery impacts what we know (perceive as fact) about it. In this paper we aim to provide an overview of the ED field’s current perspectives on recovery, discuss how our methodologies shape what is known about recovery, and suggest a broadening of our methodological “toolkits” in order to form a more complete picture of recovery. To do this, we (1) provide an overview of commonly used methodologies (quantitative, qualitative), (2) consider their benefits and limitations, (3) explore newer approaches, including mixed-methods, creative methods (e.g., Photovoice, digital storytelling), and multi-methods (e.g., quantitative, qualitative, creative methods, psycho/physiological, behavioral, laboratory, online observations), and (4) propose a potential future research model with a multi-methods design, incorporating different perspectives (e.g., increasing recruitment of diverse participants, including supportive others in study co-creation), and a longitudinal course (e.g., capturing cognitive recovery, which often comes after physical). In this way, we seek to expand our picture of what is possible in recovery; the multiple aspects of an individual’s life that can improve, the greater number of people who can recover than previously believed, and the reaffirmation of hope that, even after decades, individuals can begin and continue their ED recovery process.

How we research the process of eating disorder (ED) recovery impacts what we know (perceive as fact) about this process. Traditionally, research has focused more on the “what” of recovery (e.g., establishing criteria for recovery, connecting research and clinical experiences, reaching consensus definitions) than the “how” of recovery research (e.g., timing and framing of ED recovery items and measures, type of methodologies, triangulation of perspectives). Given that our “ways of looking” are inextricably tied to what we are looking at [ 1 ] it is important to step back and investigate how research methods shape what we can know about a phenomenon of interest. This exploration can offer insight into missing pieces of the analytic puzzle (e.g., the current gaps in our knowledge), and invite novel ways of researching ED recovery (e.g., incorporating different perspectives).

ED recovery research that is published in peer-reviewed journals most frequently uses quantitative (numerical, “objective”) Footnote 1 statistical methods, or qualitative (descriptive, “subjective”) interview methods, in order to convey their findings. In this paper, we provide an overview of commonly used methods and outline key analytic features of various types of analyses that fit within these broader method categories. We also present an examination of these commonly used methods, reflecting on the benefits and limitations of each, and what each allows us to know, or not know, about ED recovery. Following this overview, we explore mixed-methods (quantitative and qualitative), creative methods (e.g., Photovoice, digital storytelling), and multi-methods (e.g., quantitative, qualitative, creative methods, psycho/physiological, behavioral, laboratory, online observations), which may provide directions for future research, and enable new understandings of ED recovery.

Importantly, we are not suggesting that researchers abandon these commonly used quantitative or qualitative methods or that one approach is inherently better than others. Rather, we are recommending a broadening of our methodological “toolkits,” to increase the clarity of purpose of our studies, along with an alignment of the methods used. This may enable the development of more nuanced and specific insights about ED recoveries that take into account context, varied perspectives, and different positionalities. In this paper we thus aim to provide an overview of the ED field’s current perspectives on recovery, illuminate how those perspectives are necessarily informed by our methodological choices, and recommend broadening our methodological “toolkits” in order to form a more complete picture about what can be known as possible in recovery.

Ontological and epistemological stances in ED recovery research

The goal or purpose of ED recovery research depends largely on the ontological (the “what” of research) and the epistemological (the “how” of research) stances endorsed by the researcher [ 2 , 3 ]. This is often considered to be foundational in qualitative research, but is less frequently named in quantitative approaches. We give a brief overview of these stances here, because how a researcher views the world will inevitably impact the goal of the research and the methodological approaches used. In this way, we cannot present a discussion of recovery methodologies without also considering ontology and epistemology.

Ontological stance refers to what we believe can be known [ 4 , 5 ]. From a realist perspective, there is a single objective truth which exists [ 6 ]. On the other side, a relativist perspective suggests that there is no singular reality outside of human practices [ 7 ]. Researchers can therefore vary along this spectrum in terms of their assumptions about what knowledge exists.

Epistemology refers to how we can come to know this information [ 8 ]. For example, positivism argues that we can come to understand or know an objective reality through rigorous scientific practices [ 9 , 10 ]. This is the foundation of the scientific approach and what has often been referred to as the “hard sciences.” Recently, there has been a shift in which individuals from this perspective acknowledge that data collection and interpretation may be imperfect and influenced by researcher characteristics; what is now known as post-positivism [ 11 ].

Many of the quantitative approaches that will be discussed in this paper come from a post-positivist framework in that they assume there is an objective recovery “truth” that can be uncovered if we are rigorous in our approach and seek to minimize bias. On the other side, there are contextualist [ 12 ] and constructionist [ 13 ] epistemologies. Contextualism situates knowledge and the people who create it (e.g., participants, researchers) in a broader context, acknowledging that no one person can know everything. Constructionism argues that meaning is multiple, socially-constructed, and connected to wider systems of power. In this way, there is no one definition or understanding of a phenomenon.

The ontological/epistemological stance and research assumptions that dictate the approaches we take in turn inform debates on recovery. Those coming from different traditions will thus have different views of what can be known about the phenomenon. For example, the frames of (post)positivism typically underlie quantitative research, and researchers coming from this perspective have long been calling for a clear, consistent, and applicable definition of recovery (e.g., [ 14 , 15 , 16 , 17 ]). However, no overall consensus definition has been reached to date, which has several implications from a (post)positivist perspective. This lack of conceptual clarity, and between-study differences in measurement approaches, impact our ability to compare the findings between studies, including reported recovery rates, which can vary dramatically, depending upon the definitions and clinical groups used (e.g., [ 14 , 15 , 18 , 19 ]).

The belief that there is a need for a singular definition is one way of understanding the utility of recovery and may be useful for some groups. The intent of our paper, though, is not to provide a statement about what a consensus definition might be. Rather, we are offering a more diverse view of methodological perspectives (which stem from various ontological and epistemological stances) and ideas that might allow for forward movement in the field. In a dialectal format, this can involve both movement toward and away from consensus, including perspectives which do not seek to identify a single recovery definition. These paths are sometimes polarized, indicating that research aiming for (provisional) consensus is incompatible with research pushing into new areas. We suggest that both can be simultaneously pursued, acknowledging that one does not discount the other.

ED recovery research approaches: A brief overview

Quantitative research stemming from (post)positivist perspectives has tended to emphasize “objective” illness and recovery criteria that can be measured and compared in the lab/treatment, such as body mass index (BMI) (e.g., [ 20 ]), and behavioral/cognitive symptoms (e.g., [ 21 , 22 ]). For example, scores of validated ED measures such as the Eating Attitudes Test (EAT) [ 23 ] are frequently subdivided into “threshold” (criteria met for a probable clinical diagnosis) or “subthreshold” (diagnostic criteria unlikely to be met). Changes in measurable physical, behavioral, and symptomatic criteria are characteristic of the medical model of recovery, with a growing body of research suggesting that such approaches may not fit as well with lived experience perspectives [ 24 , 25 ].

In a systematic review of 126 studies looking at predictors of ED outcomes [ 26 ], symptom remission was used as a key outcome in over 80% of studies. This may differ from the “process” recovery criteria typically used in clinical settings, where the individual’s progress in therapy (e.g., how they navigate their recovery, showing improvements in not only symptoms, but also psychosocial functioning) may affect the extent to which they are deemed “recovered.”

A key element of these different definitions hinges on the extent to which symptom remission is considered an important first step in recovery. This point has often been promoted as self-evident, but is inconsistent with some orientations to recovery. For instance, a recovery model orientation, which has been noted to be potentially resonant with EDs (e.g., [ 27 ]) starts with an emphasis on a person’s goals and contexts, rather than assuming that symptom remission is a first step. This does not mean that “anything goes"; a recovery model promotes collaboration and discussion in exploring what recovery means and does for the person seeking it [ 28 , 29 ].

More recently, researchers have suggested that in alignment with this recovery model, it may be possible to continue to exhibit some symptoms (e.g., behaviors), but have improvement in other areas (e.g., improved psychosocial functioning, QOL), and still feel that one is in ED recovery (e.g., [ 27 , 29 , 30 , 31 , 32 , 33 , 34 , 35 ]). ED advocacy groups led by people with lived experience are also beginning to support approaches to harm reduction within ED recovery circles, such as those promoted by Nalgona Positivity Pride. Indeed, there are now several ED-specific, standardized measures of functioning and QOL that can provide more insight in this area, including the Eating Disorders Quality of Life (EDQOL) [ 36 ], Quality of Life for Eating Disorders (QOL ED) [ 37 ], Health-Related Quality of Life in Eating Disorders Questionnaire (HeRQoLED) [ 38 ], Eating Disorders Quality of Life Scale (EDQLS) [ 39 ], and, more recently, the Eating Disorders Recovery Questionnaire (EDRQ) [ 40 ]. In addition to assessing QOL from a quantitative perspective, QOL can also be explored qualitatively, allowing it to be contextualized against the landscape of participants’ lives.

What is considered to “matter” in ED recovery definitions thus far also differs to some extent according to who is asked. Researchers, clinicians, and people with lived experience (individuals and supportive others such as parents, family, partners, friends, mentors) may emphasize different criteria for ED recovery [ 41 ]. Further, these categories are not distinct; people may simultaneously occupy multiple positionalities at once, such as researchers and/or clinicians who also have lived experience of ED. While a consensus definition among clinicians is arguably becoming a more plausible goal [ 16 ], there may still be significant divergence of opinion between the larger clinical, research, and lived experience spaces [ 18 ]. Nevertheless, Bachner-Melman et al. [ 42 ] also found a broad area of overlap in the perspectives of people with lived experience of an ED, family members, and ED therapists, on what recovery encompasses. They proposed a questionnaire to measure four aspects of ED recovery that were agreed on by these overlapping perspectives: lack of symptoms, acceptance of self and body, social and emotional connection, and physical health [ 20 ].

Lived experience, including personally having lived with an ED, as well as being a “support” for someone with an ED (e.g., parents, family, partners, friends, mentors), necessarily informs a particular person’s ED recovery definition and provides an additional lens on the same construct. Thus, individuals who have lived through an ED may have a different view of recovery to that of their “supports” (e.g., improved psychosocial functioning, QOL, vs. medical stability, decreases in behaviors, and vice versa). However, the recovery priorities of individuals and “supports” may also align. For example, recent studies have indicated that both individuals and parents/families place high value on increased body acceptance and independence in the individual’s recovery process [ 43 , 44 ]. In addition, there is a relatively new resource for partners of those with ED, which focuses on understanding, supporting, and connecting with the partner on shared recovery goals [ 45 ]. In the book “Loving Someone with an Eating Disorder,” Dana Harron includes perspective-taking exercises to help the person understand their partner’s struggle, strategies for dealing with mealtime challenges, up-to-date facts about EDs, and self-care tips to help the person maintain healthy boundaries [ 45 ].

Conceptual and methodological challenges in ED recovery research

Related to the above, researchers face a number of conceptual and methodological questions when exploring ED recovery. For example, whether or not recovery should be considered per individual ED, or trans-diagnostically, is an important question in the recovery definition literature. Bardone-Cone et al. [ 15 ] suggest that a transdiagnostic approach is most appropriate, given that diagnoses can shift, and symptoms can fluctuate over time. Indeed, longitudinal studies have indicated that participants report receiving a single ED diagnosis at one point in time, however, over their lifetime, they would have met criteria for two, three, or four “different” ED diagnoses at different times (e.g., [ 46 ]). Anecdotally, our co-authors have also noted this when recruiting participants for research. Instead of necessitating an overall cessation of ED symptoms within one diagnostic category in the traditional categorical approach, a transdiagnostic approach could rather focus on improving the status of individual symptoms (e.g., frequency of restricting, binging) as a marker of individual “recovery.”

Beyond differences in being able to compare clinical groups across research findings, we might also consider who is most commonly included in these recovery studies (and who is not). There are many significant logistical barriers to receiving an ED diagnosis and related treatment worldwide. Indeed, practical barriers include: cost, insurance coverage, rurality, transportation, work or education schedules, and lack of available childcare, which disproportionately affects people from potentially disadvantaged groups (e.g., [ 47 , 48 , 49 , 50 ]). The process of recovery itself invokes privilege (e.g., who is able to be diagnosed, who has access to formal treatment, and who is recovering in the “right way”). For example, EDs may be missed, or diagnoses delayed, in those who do not fit the stereotypical picture of a person with an ED, including those in larger, or non-emaciated, bodies [ 51 , 52 ].

The majority of the studies thus far on ED recovery definitions are therefore composed predominantly of non-diverse participant samples who have the means to overcome the barriers to treatment access (i.e., predominantly White, thin, socioeconomically privileged, cisgender women, drawn primarily from clinical settings). Indeed, there has been comparatively little research on other populations with EDs (e.g., cis men, trans and nonbinary people, children, elders, higher weight individuals, individuals with binge eating disorder [BED], comorbidities, or late onset), as these groups often do not have access to the diagnoses and treatments that are the gateway to research study participation. These limitations determine whose recoveries we can learn about, and excludes other experiences [ 18 , 53 , 54 ].

Traditionally, those with lived experience have not been invited to co-design recovery research, limiting study participation and the diversity of representation. Even when recovery research includes non-clinical samples, methodology choices impact who is selected for participation. For example, studies that exclude potential participants with BMIs above certain levels (e.g., BMIs that are considered “overweight” or “obese”) exclude many ED recovery experiences automatically, limiting the view of what “recovery” looks like.

Additionally, the specific terminology of “recovery” may not resonate with all people experiencing life beyond an ED [ 55 ], causing some potential participants to self-select out of such studies. Some people with lived experience note that the term “recovery” is prescribed and carries preconceptions [ 56 , 57 ]. Indeed, there are nuances and connotations involved with the use of the word “recovery.” Some individuals may consider themselves “in recovery” (on a continuous journey), while others may consider themselves “recovered” (having moved past the ED completely). In this way, the meaning of “recovery” can indicate both a process and a state [ 58 ]. Stringent criteria for including people in studies as “recovered” may pre-define the group with whom recovery is being explored. Other terminology, such as severe and enduring anorexia nervosa (SEAN), and severe and enduring eating disorders (SEED), emphasize more chronic conditions. However, these terms are not always helpful for people experiencing longer-lasting ED, as they may insinuate that healthcare providers (or the patient) have given up hope for recovery [ 59 , 60 , 61 , 62 ].

Given that “recovery” as a term does not resonate with all [ 55 , 63 , 64 , 65 ], using other terms, including non-clinical ones (e.g., “getting better”, “healing”) to refer to these experiences may increase the diversity of experiences in the literature. As we will explain, these methodological features matter in recovery research because they significantly impact what we can know about ED recovery, and for whom.

Positioning ourselves

We come into this work from various vantage points; we name our positionalities here, since researchers’ subjectivity inevitably shapes their research and interpretations [ 66 ]. Engaging with the subjective, rather than presuming objectivity is the most ethical and effective stance in research, and can invite opportunities to uncover new and different knowledge [ 67 ]. The authors bring research and clinical lenses to bear on this work; some of us are primarily or exclusively researchers in the ED field, whereas others also practice clinically. We come from Global North countries, and all of us are White. We were thus trained in scientific traditions that privilege certain ways of knowing and doing that reflect the English and White dominant landscape of academia. While most of us benefit from thin privilege, able-bodied privilege, and cis-hetero privilege, our authorship team also includes those with non-binary, queer, fat, and chronically ill identities. Some of us have lived experience with ED, and have used this to inform our research and clinical practice. Some of us are newer to the ED field, whereas others have been working in the field for over 30 years. While we are different in some ways, our sameness centers around the academic privilege we have to access, interpret and navigate these literatures and their methodologies.

Overview and analysis of ED research methods

Below we provide an overview of the commonly used quantitative and qualitative research methods, along with tables that illustrate examples of the different types of analyses that fit within these broader methodological categories. We also analyze the benefits and limitations of each method, focusing on what we can learn from them and identifying relevant gaps in the literature.

Quantitative methods

As noted above, quantitative methods typically stem from a (post)positivist ontological/epistemological stance, which inherently affects how data are interpreted and understood. This is a core consideration of how we in turn can view the findings. This approach aims to provide “objective” results [ 68 ]; in this case, it is “recovery by the numbers.” It allows for the measurement of results through data, relying on a systematic approach of empirical investigation, and based on the assumption that there is a singular recovery definition which can be known. Researchers use statistical models, computational techniques, and mathematics to develop and test specific hypotheses. The types of quantitative analyses range from relatively simple descriptive/comparative measures to more complex multivariate measures and multi-level designs (which are all influenced by their study samples, assessments, and testable hypotheses). Data can be collected from the traditional in-person research study (or through video conferencing), or alternatively, from participant surveys (e.g., online, phone, mail, text).

Different types of quantitative methods have been employed in ED recovery research (see Table  1 ). Descriptive studies focus on the “how/what/when/where,” rather than the “why” (e.g., examining aspects of recovery definitions [ 69 ]), and comparative studies have a procedure to conclude that one variable is better than another (e.g., comparing different recovery definitions for agreement [ 70 ]). Univariate analyses examine the statistical characteristics of a single variable (e.g., dichotomous yes/no variable differences between recovery groups on a single measure [ 33 ], continuous range variable differences between recovery groups on multiple measures [ 71 ]), while bivariate analyses determine the empirical relationship between two variables (X and Y) (e.g., relationships between recovery attitudes and related variables [ 72 ]). Multivariate analyses aim to determine the best combination of all possible variables to test the study hypothesis (e.g., comparing recovery and healthy control groups across different recovery scores [ 73 ]).

According to (post)positivist stances, these quantitative methods have the anticipated or theoretical benefits of enabling researchers to reach higher sample sizes (increases generalizability), randomize participants (reduces bias), and replicate results (validates data). In practice, though, generalizability extends only to the sample that is recruited (as noted above, in most cases, thin, White women), and randomization within that sample thus does not increase the diversity of results. The relative focus on “novel” research means that replication studies are not conducted to the degree we would hope or expect.

In addition, quantitative methods limit what can be known about any particular individual. For example, numbers can tell us a person’s standardized assessment scores, but they do not include the detailed descriptions of the individual’s experiences. They also reduce recovery to a single experience which may overlook the tremendous diversity in lived experiences. Similarly, while statistical analyses can account for contextual confounding variables, they cannot tell us the broader factors which influence the delivery and the function of interventions.

Qualitative methods

Overall, qualitative methods offer the potential to engage deeply with phenomenon of interest, often stemming from non-positivist epistemological stances (e.g., constructionist, feminist). While qualitative methods are commonly critiqued for small sample sizes, in a qualitative paradigm, small samples allow researchers to dig into the nuances illustrated in participants’ stories, strengthening study findings. The aim of qualitative research is in-depth, contextualized analysis, rather than generalizations. Qualitative methods often, but not always, involve interacting directly with participants in the form of interviews or focus groups. However, qualitative research can also involve analyses of existing textual or image data, such as blog or social media posts, or news articles. A core feature of qualitative research is the researchers’ focus on exploring meaning in voiced or textual data, vs. using only quantitative measures.

Despite shared features, qualitative methods vary enormously in terms of data collection and analysis types. This is due in part to the differences in theoretical basis, epistemologies, ontologies, and paradigms that inform what meaning researchers perceive as possible to achieve. Some (e.g., [ 74 ]) draw on the concepts of “big Q” and “small q” to differentiate in broad terms between the qualitative methods [ 75 ]. Briefly, “big Q” methods invite and acknowledge researcher subjectivity, whereas “small q” approaches attempt to aim at more “objective interpretation,” [ 76 ], which is more similar to (post)positivist approaches. Further, “big Q” approaches tend to delve into the connections between knowledge production, analysis, and sociocultural contexts in which research takes place, whereas “small q” approaches tend to focus more on descriptive, groundwork-laying analysis for quantitative methods to provide generalizability [ 76 ]. Neither approach is inherently “better;” they are designed to achieve distinct findings.

Some of the qualitative methods that have been commonly used to explore ED recovery experiences are summarized in Table  2 . Note that these are not the only methods used. Some (particularly earlier) studies, describe their methods as “qualitative,” without specifying the exact type(s) of analysis. The differences between these various types of methods are at times subtle.

Discourse Analysis (DA) focuses on language not as just a route to content, but as powerful in and of itself (e.g., analysis of talk about recovery) [ 77 , 78 ]. Within DA, Linguistic Analysis adds a focus on language present in the text, with more of an emphasis on terms used, and their connotations (e.g., explorations of Internet message board communications about recovery) [ 79 ]. Also within DA, Narrative-Discursive Analysis adds a focus on social power (e.g., analysis of recovery interviews with a gender lens [ 80 ]), alongside an emphasis on stories (individual and broader, social stories). Narrative Approaches emphasize the story (e.g., analyses of participant writing, life-history), and situates recovery within the broader culture [ 31 , 81 , 82 , 83 , 84 ]). Phenomenological and Phenomenographic Approaches, including Interpretive Phenomenological Analysis (IPA) aim to get in “close” to participant embodied experiences (e.g., focus on recovery self-process in specific groups such as men, former patients, people in recovery from AN specifically [ 85 , 86 , 87 ]). Grounded Theory emphasizes context-specific “ground-up” theory developed from participant responses (e.g., development of cyclical, phase, and process models of recovery in/outside of treatment contexts [ 65 , 88 , 89 , 90 , 91 ]). Thematic Analysis (TA) is aimed at developing patterns/themes based on data (e.g., exploring patterns in experiences of recovery [ 92 , 93 , 94 ]) and can look quite different depending on the type of thematic analysis employed, ranging from more descriptive to more analytical. Content Analysis summarizes and organizes experiences amongst a particular group in a particular context (e.g., describing the content of interviews with specific groups, for example, athletes, or exploring the content of a particular stage of recovery, such as late-stage recovery [ 95 , 96 , 97 , 98 , 99 ]).

It is also possible to use different qualitative methods to explore similar phenomenon. For example, a constructivist grounded theory exploration of people who have received treatment for AN may focus on theorizing what ED recovery processes are occurring for this particular group [ 99 , 100 ]. An IPA of this same group, meanwhile, may emphasize the development of a set of themes relating to shared perspectives on what the experience felt like [ 101 , 102 ].

Qualitative methods can thus provide detailed descriptions of a wide diversity of lived experiences. This enables us to have a broader perspective of what is possible in recovery. Additionally, these methods allow us to consider the contextual factors which influence the delivery and the function of interventions. A potential further benefit is the individual’s own process of reflecting on their changes in recovery (via study participation), which may provide insight and encouragement for continuing on their path.

Critiques of qualitative methods tend to center around the concept of generalizability, though as noted this is not typically the goal of qualitative approaches. As noted above, quantitative studies, which typically focus on a person’s ED standardized assessment scores, and account for contextual confounding variables through statistical analyses, theoretically generate findings which can be applied to other populations from the study sample. However, as we have indicated, extrapolation of the results of mostly homogenous groups (e.g., predominantly White, thin, socioeconomically privileged, cisgender women, drawn primarily from clinical settings) falsely assumes that the course and outcomes will be the same for all.

Exploration of mixed-methods, creative methods, and multi-methods research

While ED recovery researchers have primarily conducted either quantitative or qualitative studies, some have integrated alternative or multiple methods in their designs. Below we explore some of these methods, which may enable new understandings of ED recovery. These include mixed-methods (usually a weaving of quantitative and qualitative), creative methods (e.g., Photovoice, digital storytelling), and multi-methods (e.g., complementary combinations of quantitative, qualitative, creative methods, psycho/physiological, behavioral, laboratory, online observations).

Mixed-methods (weaving of quantitative and qualitative)

It has been argued that mixed-methods allow us to weave together our quantitative and qualitative insights by acknowledging the benefits and limits of both designs. Cluster analysis intends to combine these methods with the goal of maximizing benefits [ 103 ]. For the qualitative aspects of the analysis, data is coded to themes using one of the qualitative methods, and each individual unit of data (e.g., person) is coded for the presence or absence of each theme. For the quantitative aspects of the analysis, data is plotted to identify different clusters of individuals, and then these clusters are interpreted via statistical methods. Cluster analysis aims to provide greater insight into groups of individuals, and potentially elucidate different clusters of “recovery definitions.” However, the qualitative analysis in this approach is inherently reductionistic (e.g., people are coded to create a quantitative measure), which aligns with the post-positivist stance associated with quantitative analyses. From this view, the approaches are not actually integrated; rather they are complimentary. Indeed, it may not be possible to truly integrate them when they emerge from different epistemological stances. We suggest, however, that integration is not needed.

Bachner-Melman et al. [ 42 ] used exploratory factor analysis to identify four factors that mapped onto ED recovery which had general agreement between participants with a lifetime ED diagnosis, healthy family members, and ED clinicians; (1) lack of symptomatic behavior, (2) acceptance of self and body, (3) social and emotional connection, and (4) physical health. These factors were then confirmed using confirmatory factor analysis. Utilizing more than one method thus expands our perspectives of ED recovery, allowing us to broaden our understanding of what can be known about recovery. Yet as noted above, we caution readers in viewing mixed-method approaches as an overall panacea; the approach tends to be more (post)positivist, and aims to quantify experiences, which may not be the goal for researchers from other stances. Again, this is not to say that the approach is without merit, but that it is important to acknowledge what it aims to do (or know).

Creative methods (photovoice, digital/verbal storytelling, collages, drawings)

Quantitative and qualitative methods are, of course, not the only options at the disposal of researchers interested in exploring ED recovery. Some researchers have elected to take creative approaches to research, seeking to explore recovery in different ways. Potentially, such methods enable researchers to “see” facets of recovery phenomenon that are less evident in methods that primarily hinge on either words or numbers [ 57 ]. To date, ED recovery researchers have used creative methods such as Photovoice [ 104 , 105 ], which aims to involve participants in the process of generating and analyzing research data [ 106 ]. This method may be particularly useful for generating disseminable results, with a view towards change in policy settings for the benefit of people in recovery [ 104 ].

Another creative method, digital storytelling [ 57 , 107 ], encourages participants to “story themselves” at a particular moment in time. This may enable the creation of more nuanced, rich, and person-centered depictions of recovery; the participant’s voice is centered in a way that may be less feasible in research that seeks to generate patterns across several participants’ accounts [ 107 ]. Like Photovoice, digital stories can also be used to work toward enhancing understandings of recovery amongst people who do not have lived experience (e.g., healthcare providers) [ 57 ].

Other creative methods include the use of collages, verbal storytelling, drawing,and more [ 108 , 109 , 110 ]. Placing the decision about which creative method to use in the hands of research participants may also enable a redressing of traditional power dynamics in research that position the researcher as the ultimate decision-maker [ 57 ].

Multi-methods (complimentary use of multiple methods)

Given that ED recovery is a complex phenomenon, one approach to exploring it is the use of a multi-method research design, including different complimentary types of analyses (e.g., quantitative, qualitative, creative methods, psycho/physiological, behavioral, laboratory, online observations), and ideally from different perspectives (e.g., individuals, “supports,” clinicians, researchers, policy makers, and other stakeholders). Broadening our methodological “toolkits” may allow for more nuanced and specific insights about ED recoveries, taking into account context, varied perspectives, and positionalities.

Several studies of ED symptom assessment have employed multi-method designs thus far, and each of these has the potential to contribute a piece of the “ED recovery puzzle.” For example, Stewart et al. [ 111 ] conducted a mixed method investigation of the experiences young people, parents, and clinicians had of online ED treatment during COVID-19. They used a mixed quantitative (Likert scale rating questions) and qualitative (free text entry questions) survey which they analyzed using a summary approach (quantitative) and reflexive thematic analysis (qualitative). Leehr and colleagues [ 112 ] analyzed binge eating episodes under negative mood conditions via electroencephalography (EEG) and eye tracking (ET) in a laboratory.

Bartholome et al. [ 113 ] combined standardized instrument interviews, laboratory investigations, and ecological momentary assessment (EMA) to collect data on binge eating episodes in participants with BED. In order to examine underlying mechanisms of the somatic sensation of “feeling fat,” Mehak and Racine [ 114 ] used multiple methods of self-reports, EMA, heart rate variability, laboratory measurements of BMI, dual energy X-ray absorptiometry (DEXA) scans, and clothing sizes.

Technological advances can offer more in-depth information about the recovery process, and could be utilized further in research studies. For example, EMA allows for moment by moment collection of data on phenomenon of interest. This can include biological functions (e.g., heart rate), as well as feelings and behaviors in an individual’s daily life. This approach provides a more “real time” look into experiences (versus having to recall such experiences later on in a survey, or in an interview). In the case of ED recovery, future research can explore the answer to the question of what “active” recovery looks like on a daily basis for individuals via EMA (e.g., what challenges do individuals encounter; how does it impact their behavior/feelings?).

Other technological laboratory tools, such as fMRI, DEXA, and other scanning techniques, can add visual information about the current physical state of recovery (and any related functional “scars” from the ED). It may be helpful in future studies to provide this feedback so individuals can have an accurate picture of the medical status of their body, and make adjustments (e.g., take Vitamin D to increase bone density).

Additional methodological approaches/considerations

Co-design with different perspectives in ed recovery research.

We believe it will be helpful to incorporate different perspectives in ED recovery research for many reasons. For example, expanding who is included in our studies, what identities are represented, and those with both formal and informal treatment, will increase the participant representation, relevance, and generalizability of the findings [ 53 ]. This in turn will allow us to know more about the process of recovery for different individuals and groups, and will broaden our conceptualization of the phenomenon. Ideally, future research will include individuals with lived experiences and their “supports” (e.g., parents, families, partners, friends, mentors), as well as clinicians and researchers, in co-designing studies that could identify and assess aspects of ED recovery that are important to all of the constituents.

Longitudinal research design

We also underscore the need for an extended duration of studies in order to better understand the longitudinal course and outcome of individuals in ED recovery. This design will allow us to compare recovery operalizations vs. subsequent relapse rates, to track how perspectives of recovery develop and change over time (via quantitative, qualitative and mixed-method measures), and how these in turn affect an individual’s identity (e.g., [ 15 ]).

There are several areas of potential future longitudinal research. For example, follow-up on cognitive recovery (which we know tends to occur later (e.g., [ 43 ]), and how related timelines for this may impact subsequent relapse rates, tracking recovery changes over time with mixed or multi-methods designs in underrepresented populations (e.g., Atypical Anorexia (AAN) [ 115 ]), and holding on to hope, with more longitudinal data indicating that recovery is possible, even after decades (e.g., [ 59 ]).

Future research directions

Based on the above studies, we suggest some potential areas for future research, ideally incorporating multi-method designs to provide different perspectives on ED recovery. Recently, several themes have been identified in the literature as promising lines of research that may improve our understanding of ED, and increase the clinical application of findings.

Predictors of outcomes

Within quantitative research, Bardone-Cone et al. [ 15 ] note that predictors of outcomes, biological/neuropsychological techniques, and a focus on the SEAN population are newer, more nuanced, areas of investigation. In their systematic review and meta-analysis of predictors of ED treatment outcomes (at end of treatment [EoT], and follow-up), Vall and Wade [ 26 ] reported that the most robust predictor at both time frames was greater symptom change earlier in treatment. Other baseline predictors of better outcomes included: higher BMI, fewer binge/purge behaviors, more functional relationships (e.g., with family, friends), and greater motivation to recover. Of note, it is important to understand that higher BMI is in the context of a “higher” thin BMI, as most people with BMI > 25 are not included in studies of recovery. This is another example of how our methods and design choices impact what we can know.

Relatedly, one potential area for future ED quantitative predictor research is to build a Risk Calculator (RC), which is a statistical tool that identifies risk factors, and determines how likely an event is to happen for a particular person [ 116 ]. Physicians have used RCs clinically across an array of medical conditions, including stroke [ 117 ] and cancer [ 118 ]. There has been a recent turn toward integrating RCs into charting psychiatric disorder outcomes and treatment approaches; they have been used for psychosis [ 116 ], depression [ 119 ], and bipolar disorder [ 120 , 121 , 122 , 123 ]. This same technique could be applied to build a RC for personalized risk of ED onset/relapse, utilizing variables collected in research/treatment. A statistical combination of factors that reliably predict the non-occurrence of ED relapse could be a valuable addition to, predictor of, or even criterion, for full recovery. As part of these research initiatives, it will be important for researchers to employ diverse and longitudinal methods in order to obtain long-term, dynamic data.

In line with this, narrative qualitative analysis may be useful in elucidating predictors/risk factors for individuals. In this way, the (narrative) story that the person tells themselves about their recovery, and what was helpful to them, also has importance alongside any quantitative measures. Indeed, this perspective perhaps has more personal meaning, especially in contexts where minute changes identifiable through quantitative studies may be less relevant in the daily lives of their ED recovery.

Biological and neurological markers

Recent developments in the understanding of biological and neurological markers have enabled us to parse out what features may be involved with the ED “state” (which resolves with recovery), what features may onset premorbid to the ED (and will potentially continue after recovery), and what features may be “scars” (consequences of the ED). In their functional Magnetic Resonance Imaging (fMRI) study on participants who had recovered from AN, Fuglset et al. [ 124 ] reported increased activation in visual processing regions in anticipation of seeing images of food, with corresponding reduced activation in decision-making regions. While they found some normalization of the brain regions during recovery, other differences related to longer periods of starvation that appear later in life remained (residual “scars”).

Future quantitative research which incorporates longitudinal designs following the same participant cohort may elucidate more closely the timepoints during which the “state” and “scar” markers begin to emerge, in order to provide earlier interventions. In this instance, qualitative longitudinal studies could be beneficial here too. For example, participant narrative descriptions of ongoing biological changes in their recovery (e.g., feeling hungrier, not being able to tolerate hunger as well) not only reaffirm that people are noticing these internal bio markers, but provide the opportunity for them to discuss their day to day experiences of these lasting changes.

Recovery criteria

From a post-positivist perspective, there has been a longstanding call for standardized ED recovery criteria, typically involving weight, behavioral, and cognitive criteria (e.g., [ 14 , 15 , 16 , 17 ]). Drawing from the above, we suggest that research looking into recovery criteria may benefit from more diverse methodological approaches which pull from a variety of sources (e.g., clinicians, researchers, individuals with lived experience). For example, BMI has historically been used as an indicator of recovery status because it is readily obtained by ED researchers (and clinicians). While weight monitoring can be helpful in specific cases (e.g., those who are severely underweight or have lost a lot of weight in a short period of time), it is limited in use. Namely, BMI is insufficient to determine medical stabilization, it fails to take into account individual differences, and it can have negative impacts on treatment when individuals are discharged on the basis of weight alone [ 125 ]. Given these concerns, future research could discontinue the use of BMI as the “core” recovery criterion, as suggested by Kenny and Lewis [ 126 ], and instead focus on other variables that are more indicative of recovery over follow-up (e.g., Vall and Wade’s systematic review and meta-analysis findings of early symptom change during treatment as the most robust predictor of outcomes) [ 26 ].

Similarly, standardized assessments (e.g., EDE-Q, EDE, ED-LIFE) have been the “go-to” for assessing recovery outcomes in comparison to statistical norms. However, these measures are often developed by clinicians/researchers (thus reflecting what they feel is important in recovery) and in line with particular therapeutic modalities (e.g., the EDE-Q has a cognitive orientation). Thus, scores on these measures may not always match the person’s particular recovery aims and goals, nor the relative importance of particular behaviors in their lives. Employing these measures as a part of multi-methods designs with other types of assessments for comparison may offer the potential to think differently about these measures, and their role in assessing outcomes. We also suggest the need for measures co-designed with folks with lived experience and which reflect the diverse recovery elements described in qualitative studies (e.g., [ 33 ]).

The recovery process

Future research could compile more comprehensive lived experience narratives of changes in thought patterns through the recovery journeys (e.g., descriptions of how the “ED” voice began to leave, if ED voice is a relevant construct for the person), which could provide a more realistic timeline of this portion of the process for individuals and their “supports.” To begin to employ these kinds of measures in a way that opens up new possibilities, it would also be important to explore whether the ED-related ideas being measured resonate with the person whose recovery is being explored. Co-design processes may also be particularly relevant here, inviting people in recovery to be a part of research teams and take a role in determining the kinds of measures that could be used to assess recovery.

As years of ED behaviors and thoughts tend to impair different areas of psychosocial functioning (e.g., relationships, school/work, recreation, household duties), improvement in these areas, along with related QOL, tends to also lag behind physical recovery (e.g., [ 43 ]). Future research could further elaborate the timelines for which recovery in the different areas occurs, both from a group (e.g., through life story, narrative, or thematic analysis), and an individual (e.g., personal recording of recovery progress, case study approach) level. This approach offers a shift in methodological perspective, providing opportunity to conceptualize recovery differently.

Other areas for future consideration

Several studies (and informal support groups) have successfully employed recovered mentors, providing hope in recovery (e.g., [ 127 ]). Future research could examine more of the nuances of the mentorship role, including the characteristics of the mentor, the stage of recovery that the individual is in, and the dynamics of the mentor relationship. Further, taking a truly co-designed approach and, in particular, working with those who have not been included and heard in either treatment or research (not only more diverse participants, but their “supports,” including mentors), could offer new insight into recovery processes. Indeed, going forwards, we need to conduct our research differently if we want to incorporate the perspectives of those that we do not usually hear from.

Another area of study has developed around online (e.g., social media) use among those who are at risk for developing an ED, struggling with an ED, and those who are on the path to ED recovery. Analyses of online websites, blogs, and social media posts, along with their related potentially triggering content, have been conducted (e.g., [ 128 ]). However, on a positive note, this medium allows us to explore other methodological possibilities, including potentially focusing on reducing participant burden, and engaging with content from spaces where people are more “organically” describing these experiences, to get a sense of recovery outside of a clinical perspective. One possibility for future research is to combine the use of EMA with exposure to a range of different ED blog content (e.g., from triggering to supportive posts), in order to provide more proximal individual reaction information (e.g., EMA before exposure, EMA at exposure time, EMA after exposure time).

Proposed future research model: Dialectical movement towards and away from a consensus

The aim of this paper is to offer a more diverse view of methodological perspectives (which stem from various ontological and epistemological stances) and ideas that might allow for forward movement in the field. As noted above, in a dialectal format, this can involve both movement toward and away from a consensus, including perspectives which do not seek to identify a single recovery definition. We believe that both can be simultaneously pursued, acknowledging that one does not discount the other. We have outlined several potential areas to explore which do not necessarily depend upon a consensus definition. Here, for balance, we would like to propose a future research model that could guide us in a direction that may eventually lead to a consensus definition–or definition s . In effect, we are advocating for: (1) transparency in researchers’ epistemological stances; (2) more varied approaches to research in order to “see” different aspects of recovery experiences; and (3) collaboration between researchers and other stakeholders to generate new methodological approaches and insights about recovery.

Our proposed future research model is detailed in Table  3 . Based upon the studies we cited above, we suggest a multi-methods design (e.g., quantitative, qualitative, creative methods, psycho/physiological, behavioral, laboratory, online observations), which incorporates different perspectives (e.g., expanding recruitment of participants that have been less represented in the literature, including “supportive” others), and extends the duration of studies to provide a more longitudinal outlook (e.g., capturing cognitive recovery, and improvement in psychosocial functioning/QOL, which often comes later, and noting how definitions of recovery may change over time for people). In this way, we hope to move the field towards different, more nuanced, and comprehensive perspectives on ED recovery.

In conclusion, we would like to encourage a creative, transparent, and thoughtful approach to ED recovery methodology, that considers what each of the methods allows us to engage with, or not, as the case may be. What we can (and do) know about recovery is intricately tied to our methodological and study design choices, which all have limits. Within this context, while there is a benefit to current pushes in the field to “come to consensus,“ these consensus definitions will necessarily leave out some people and experiences. This is especially the case for those who have not been meaningfully included in the research we have conducted to reach this consensus (e.g., people with lived experience, “non-traditional” patients, patients without access to treatment). Since there are so many different facets of recovery experiences, using different methodologies is imperative to develop a more complete understanding.

Indeed, it is important to acknowledge how the centrality of the method that is chosen to define ED recovery in turn influences how researchers and clinicians understand recovery, and how one moves towards it. New insights into recovery processes may depend on new methods of investigation. Thus, we suggest that some potential areas for future research ideally employ multi-method designs (e.g., quantitative, qualitative, creative methods, psycho/physiological, behavioral, laboratory, online observations), incorporate different perspectives (e.g., expanding recruitment of participants that have been less represented in the literature, including supportive others) and extend the duration of studies to provide a more longitudinal outlook (e.g., capturing cognitive recovery, which often comes later, and noting how definitions of recovery may change over time for people). In this way, we hope to move the field towards different, more nuanced, and comprehensive perspectives on ED recovery.

Availability of data and materials

Not applicable.

There are debates about the degree to which research can ever be truly objective or whether this is desirable. Here, we use objective and subjective in quotation marks to signal broader perceptions about these processes.

Abbreviations

Anorexia nervosa

Binge eating disorder

Body mass index

Bulimia Nervosa

Connectedness, hope and optimism, identity, meaning in life, empowerment

Dual energy X-ray absorptiometry

Eating attitudes test

  • Eating disorders

Eating Disorders Recovery Endorsement Questionnaire

Eating Disorders Recovery Questionnaire

Eating Disorders Quality of Life Scale

Eating Disorders Quality of Life

Electroencephalography

Ecological momentary assessment

Eye tracking

Functional magnetic resonance imaging

Generalized estimating equations

Health-Related Quality of Life in Eating Disorders Questionnaire

Hierarchical linear models

Interpretive phenomenological analysis

Multivariate analysis of variance

National Institutes of Health

Risk calculator

Substance Abuse and Mental Health Services Administration

Severe and enduring anorexia nervosa

Severe and enduring eating disorder

Quality of life for eating disorders

Barad K. Meeting the Universe Halfway. Durham: Duke University Press; 2007.

Book   Google Scholar  

Slevitch L. Qualitative and quantitative methodologies compared: ontological and epistemological perspective. J Qual Assur Hosp Tour. 2011;12(1):73–81.

Article   Google Scholar  

Tuli F. The basis of distinction between qualitative and quantitative research in social science: reflection on ontological, epistemological and methodological perspectives. Ethiop J Educ Sci. 2010;6(1).

Duncan C, Cloutier JD, Bailey PH. Concept analysis: the importance of differentiating the ontological focus. J Adv Nurs. 2007;58(3):293–300.

Article   PubMed   Google Scholar  

Hofweber T. Logic and ontology. In: Zalta EN, editor. The Stanford encyclopedia of philosophy. Stanford: The Metaphysics Research Lab, Center for the Study of Language and Information; 2021.

Google Scholar  

Hupcey JE, Penrod J. Concept analysis: examining the state of the science. Res Theory Nurs Pract. 2005;19(2):197–208.

Wilson J. Thinking with concepts. Cambridge: Cambridge University Press; 1963.

Setup M, Ram N. Epistemology. In: Zalta EN, editor. The Stanford encyclopedia of philosophy. Stanford: The Metaphysics Research Lab, Center for the Study of Language and Information; 2020.

Macionis JJ, Gerber LM, Sociology (7th Canadian). Pearson, Canada: Pearson Canada; 2011.

Larrain J. The concept of ideology. London: Hutchison; 1979.

Phillips DC, Burbules NC. Postpositivism and educational research. Lanham & Boulder: Rowman & LIttlefield Publishers; 2000.

Stanley J. Knowledge and practical interests. New York: Oxford University Press; 2005.

Cakir M. Constructivist approaches to learning in science and their implications for science pedagogy: a literature review. Int J Environ Sci Educ. 2008;3(4):193–206.

Bardone-Cone AM, Harney MB, Maldonado CR, Lawson MA, Robinson DP, Smith R, et al. Defining recovery from an eating disorder: conceptualization, validation, and examination of psychosocial functioning and psychiatric comorbidity. Behav Res Ther. 2010;48(3):194–202.

Bardone-Cone AM, Hunt RA, Watson HJ. An overview of conceptualizations of eating disorder recovery, recent findings, and future directions. Curr Psychiatry Rep. 2018;20(9):79.

Wade TD, Lock J. Developing consensus on the definition of remission and recovery for research. Int J Eat Disord. 2020;53(8):1204–8.

Attia E, Blackwood KL, Guarda AS, Marcus MD, Rothman DJ. Marketing residential treatment programs for eating disorders: a call for transparency. Psychiatr serv. 2016;67(6):664–6.

McGilley BS.  J. Recipe for recovery: necessary ingredients for the client’s and clinician’s Success. In: Maine M, Bunnell, editors. New York: Elsevier; 2010.

Noordenbos G. Which criteria for recovery are relevant according to eating disorder patients and therapists? Eat Disord. 2011;19(5):441–51.

Solmi M, Wade TD, Byrne S, Del Giovane C, Fairburn CG, Ostinelli EG, et al. Comparative efficacy and acceptability of psychological interventions for the treatment of adult outpatients with anorexia nervosa: a systematic review and network meta-analysis. Lancet Psychiatry. 2021;8(3):215–24.

Treasure J, Duarte TA, Schmidt U. Eating disorders. Lancet. 2020;395(10227):899–911.

Waller G, Raykos B. Behavioral interventions in the treatment of eating disorders. Psychiatr Clin North Am. 2019;42(2):181–91.

Garner DM, Garfinkel PE. The eating attitudes test: an index of the symptoms of anorexia nervosa. Psychol Med. 1979;9(2):273–9.

Kenny TE, Trottier K, Lewis S. Lived experience perspectives on a model of eating disorder recovery in a sample of predominantly white women: a mixed method study. J Eat Disord. 2022;10(1):1–19.

McDonald S, Williams AJ, Barr P, McNamara N, Marriott M. Service user and eating disorder therapist views on anorexia nervosa recovery criteria. Psychol Psychother. 2021;94(3):721–36.

Article   PubMed   PubMed Central   Google Scholar  

Vall E, Wade TD. Predictors of treatment outcome in individuals with eating disorders: a systematic review and meta-analysis. Int J Eat Disord. 2015;48(7):946–71.

Dawson L, Rhodes P, Touyz S. “Doing the impossible”: the process of recovery from chronic anorexia nervosa. Qual Health Res. 2014;24(4):494–505.

Anthony WA. Recovery from mental illness: the guiding vision of the mental health service system in the 1990s. Psychosoc Rehabilit J. 1993;24(9):159–68.

Wetzler S, Hackmann C, Peryer G, Clayman K, Friedman D, Saffran K, et al. A framework to conceptualize personal recovery from eating disorders: a systematic review and qualitative meta-synthesis of perspectives from individuals with lived experience. Int J Eat Disord. 2020;53(8):1188–203.

Jarman M, Walsh S. Evaluating recovery from anorexia nervosa and bulimia nervosa: integrating lessons learned from research and clinical practice. Clin Psychol Rev. 1999;19(7):773–88.

Dawson L, Rhodes P, Touyz S. Defining recovery from anorexia nervosa: a Delphi study to determine expert practitioners’ views. Adv Eat Disord: Theory Res Pract. 2015;3:165–76.

’t Slof-OpLandt MCT, Dingemans AE, de la Torre YRJ, van Furth EF. Self-assessment of eating disorder recovery: absence of eating disorder psychopathology is not essential. Int J Eat Disord. 2019;52(8):956–61.

de Vos JA, Radstaak M, Bohlmeijer ET, Westerhof GJ. Having an eating disorder and still being able to flourish? Examination of pathological symptoms and well-being as two continua of mental health in a clinical sample. Front Psychol. 2018;9:2145.

Emanuelli F, Waller G, Jones-Chester M, Ostuzzi R. Recovery from disordered eating: sufferers’ and clinicians’ perspectives. Eur Eat Disord Rev. 2012;20(5):363–72.

Noordenbos G, Seubring A. Criteria for recovery from eating disorders according to patients and therapists. Eat Disord. 2006;14(1):41–54.

Engel SG, Wittrock DA, Crosby RD, Wonderlich SA, Mitchell JE, Kolotkin RL. Development and psychometric validation of an eating disorder-specific health-related quality of life instrument. Int J Eat Disord. 2006;39(1):62–71.

Abraham SF, Brown T, Boyd C, Luscombe G, Russell J. Quality of life: eating disorders. Aust N Z J Psychiatry. 2006;40(2):150–5.

Las Hayas C, Quintana JM, Padierna A, Bilbao A, Munoz P, Madrazo A, et al. The new questionnaire health-related quality of life for eating disorders showed good validity and reliability. J Clin Epidemiol. 2006;59(2):192–200.

Adair CE, Marcoux GC, Cram BS, Ewashen CJ, Chafe J, Cassin SE, et al. Development and multi-site validation of a new condition-specific quality of life measure for eating disorders. Health Qual Life Outcomes. 2007;5:23.

Bachner-Melman R, Lev-Ari L, Zohar AH, Linketsky M. The Eating Disorders Recovery Questionnaire: psychometric properties and validity. Eat Weight Disord. 2021;26(8):2633–43.

Noordenbos G. When have eating disordered patients recovered and what do the DSM-IV criteria tell about recovery? Eat Disord. 2011;19(3):234–45.

Bachner-Melman R, Lev-Ari L, Zohar AH, Lev SL. Can recovery from an eating disorder be measured? Toward a standardized questionnaire. Front Psychol. 2018;9:2456.

Accurso EC, Sim L, Muhlheim L, Lebow J. Parents know best: caregiver perspectives on eating disorder recovery. Int J Eat Disord. 2020;53(8):1252–60.

Richmond TK, Woolverton GA, Mammel K, Ornstein RM, Spalding A, Woods ER, et al. How do you define recovery? A qualitative study of patients with eating disorders, their parents, and clinicians. Int J Eat Disord. 2020;53(8):1209–18.

Harron D. Loving Someone with an Eating Disorder. Oakland: New Harbinger Publications, Inc.; 2019.

Milos G, Spindler A, Schnyder U, Fairburn CG. Instability of eating disorder diagnoses: prospective study. Br J Psychiatry. 2005;187:573–8.

Becker AE, Franko DL, Speck A, Herzog DB. Ethnicity and differential access to care for eating disorder symptoms. Int J Eat Disord. 2003;33(2):205–12.

Cachelin FM, Rebeck R, Veisel C, Striegel-Moore RH. Barriers to treatment for eating disorders among ethnically diverse women. Int J Eat Disord. 2001;30(3):269–78.

Cachelin FM, Striegel-Moore RH. Help seeking and barriers to treatment in a community sample of Mexican American and European American women with eating disorders. Int J Eat Disord. 2006;39(2):154–61.

Ali K, Farrer L, Fassnacht DB, Gulliver A, Bauer S, Griffiths KM. Perceived barriers and facilitators towards help-seeking for eating disorders: a systematic review. Int J Eat Disord. 2017;50(1):9–21.

Lebow J, Sim LA, Kransdorf LN. Prevalence of a history of overweight and obesity in adolescents with restrictive eating disorders. J Adolesc Health. 2015;56(1):19–24.

Harrop EN, Mensinger JL, Moore M, Lindhorst T. Restrictive eating disorders in higher weight persons: a systematic review of atypical anorexia nervosa prevalence and consecutive admission literature. Int J Eat Disord. 2021;54(8):1328–57.

Egbert AH, Hunt RA, Williams KL, Burke NL, Mathis KJ. Reporting racial and ethnic diversity in eating disorder research over the past 20 years. Int J Eat Disord. 2022;55(4):455–62.

Mikhail ME, Klump KL. A virtual issue highlighting eating disorders in people of black/African and indigenous heritage. Int J Eat Disord. 2021;54(3):459–67.

Conti JE. Recovering identity from anorexia nervosa: women’s constructions of their experiences of recovery from anorexia nervosa over 10 years. J Constructivist Psychol. 2018;31(1):72–94.

Kenny TE, Thomassin K, Trottier K, Lewis SP. Views on the term ‘recovery’ in a sample of primarily white treatment-experienced women with lived eating disorder experience. Int J Eat Disord. Under Review.

LaMarre A, Rice C. Healthcare providers’ engagement with eating disorder recovery narratives: opening to complexity and diversity. Med Humanit. 2020;47:78–86.

Fava GA. The concept of recovery in affective disorders. Psychother Psychosom. 1996;65(1):2–13.

Eddy KT, Tabri N, Thomas JJ, Murray HB, Keshaviah A, Hastings E, et al. Recovery from anorexia nervosa and bulimia nervosa at 22-year follow-up. J Clin Psychiatry. 2017;78(2):184–9.

Dawson L, Rhodes P, Touyz S. The recovery model and anorexia nervosa. Aust N Z J Psychiatry. 2014;48(11):1009–16.

Touyz S, Hay P. Severe and enduring anorexia nervosa (SE-AN): in search of a new paradigm. J Eat Disord. 2015;3:26.

Calugi S, Milanese C, Sartirana M, El Ghoch M, Sartori F, Geccherle E, et al. The eating disorder examination questionnaire: reliability and validity of the Italian version. Eat Weight Disord. 2017;22(3):509–14.

LaMarre A, Rice C, Rinaldi J tracing fatness through the eating disorder assemblage. Thickening fat, 1st Edition: Routledge; 2019. pp. 64–76.

Shohet MN, Anorexia. “Full” and “struggling.” Genres  Recover Ethos. 2008;35(3):344–82.

Musolino C, Warin M, Wade T, Gilchrist P. Developing shared understandings of recovery and care: a qualitative study of women with eating disorders who resist therapeutic care. J Eat Disord. 2016;4:36.

Malterud K. Qualitative research: standards, challenges, and guidelines. Lancet. 2001;358(9280):483–8.

Gough B, Madill A. Subjectivity in psychological science: from problem to prospect. Psychol Methods. 2012;17(3):374–84.

Matthews B, Ross L. Research methods. a practical guide for the social sciences. Harlow: Pearson Education; 2010.

Couturier J, Lock J. What is recovery in adolescent anorexia nervosa? Int J Eat Disord. 2006;39(7):550–5.

Ackard DM, Richter SA, Egan AM, Cronemeyer CL. What does remission tell us about women with eating disorders? Investigating applications of various remission definitions and their associations with quality of life. J Psychosom Res. 2014;76(1):12–8.

Cogley CB, Keel PK. Requiring remission of undue influence of weight and shape on self-evaluation in the definition of recovery for bulimia nervosa. Int J Eat Disord. 2003;34(2):200–10.

Dimitropoulos G, McCallum L, Colasanto M, Freeman VE, Gadalla T. The effects of stigma on recovery attitudes in people with anorexia nervosa in intensive treatment. J Nerv Ment Dis. 2016;204(5):370–80.

Bachner-Melman R, Zohar AH, Ebstein RP. An examination of cognitive versus behavioral components of recovery from anorexia nervosa. J Nerv Ment Dis. 2006;194(9):697–703.

Braun VC, Can V. I use TA? Should I use TA? Should I not use TA? Comparing reflexive thematic analysis and other pattern-based qualitative analytic approaches. Couns Psychother Res. 2021;21(1):37–47.

Kidder LH, Fine M. Qualitative and quantitative methods: when stories converge. Eval Policy. 1987;1987(35):57–75.

Clarke V. Navigating the messy swamp of qualitative research: are generic reporting standards the answer? In: Levitt HM, editor. Reporting qualitative research in psychology: how to meet APA Style journal article reporting standards, revised edition. Washington: American Psychological Association; 2020.

Hardin PK. Social and cultural considerations in recovery from anorexia nervosa: a critical poststructuralist analysis. ANS Adv Nurs Sci. 2003;26(1):5–16.

Keski-Rahkonen A, Tozzi F. The process of recovery in eating disorder sufferers’ own words: an Internet-based study. Int J Eat Disord. 2005;37(Suppl):S80-6 discussion S7-9.

Maison H, Bailey L, Clarke S, Treasure J, Anderson G, Kohn M. Un/imaginable future selves: a discourse analysis of in-patients’ talk about recovery from an eating disorder. Eur Eat Disord Rev. 2011;19(a):25–36.

Moulding N. Gendered intersubjectivities in narratives of recovery from an eating disorder. J Women Soc Work. 2016;31(1):70–83.

LaMarre A, Rice C. Normal eating is counter-cultural: embodied experiences of eating disorder recovery. J Commun Appl Soc Psychol. 2016;26:136–49.

Matusek JA, Knudson RM. Rethinking recovery from eating disorders: spiritual and political dimensions. Qual Health Res. 2009;19(5):697–707.

Patching J, Lawler J. Understanding women’s experiences of developing an ED and recovery: a life-history approach. Nurs Inq. 2009;12(1):10–21.

Redenbach J, Lawler J. Recovery from disordered eating: what life histories reveal. Contemp Nurse. 2003;15(1–2):148–56.

Bjork T, Wallin K, Pettersen G. Male experiences of life after recovery from an eating disorder. Eat Disord. 2012;20(5):460–8.

Bjork T, Ahlstrom G. The patient’s perception of having recovered from an eating disorder. Health Care Women Int. 2008;29(8):926–44.

Jenkins J, Ogden J. Becoming ‘whole’ again: a qualitative study of women’s views of recovering from anorexia nervosa. Eur Eat Disord Rev. 2012;20(1):e23–31.

D’Abundo M, Chally P. Struggling with recovery: participant perspectives on battling an eating disorder. Qual Health Res. 2004;14(8):1094–106.

Krentz A, Chew J, Arthur N. Can J Couns. 2005;39(2):118–35.

Lamoureux MMH, Bottorff JL. “Becoming the real me”: Recovering from anorexia nervosa. Heatlh Care for Women International. 2005;26:170–88.

Woods S. Untreated recovery from eating disorders. Adolescence. 2004;39(154):361–71.

PubMed   Google Scholar  

Hay PJ, Cho K. A qualitative exploration of influences on the process of recovery from personal written accounts of people with anorexia nervosa. Women Health. 2013;53(7):730–40.

LaMarre A, Rice C. Hashtag recovery: #eating disorder recovery on instagram. Soc Sci. 2017;6(3):68.

Lord VM, Reiboldt W, Gonitzke D, Parker E, Peterson C. Experiences of recovery in binge-eating disorder: a qualitative approach using online message boards. Eat Weight Disord. 2018;23(1):95–105.

Arthur-Cameselle JN, Quatromoni PA. A qualitative analysis of female college athletes’ eating disorder recovery experiences. Sport Psychol. 2014;28:334–46.

Lindgren BM, Enmark A, Bohman A, Lundstrom M. A qualitative study of young women’s experiences of recovery from bulimia nervosa. J Adv Nurs. 2015;71(4):860–9.

Nilsson K, Hagglof B. Patient perspectives of recovery in adolescent onset anorexia nervosa. Eat Disord. 2006;14(4):305–11.

Pettersen G, Thune-Larsen KB, Wynn R, Rosenvinge JH. Eating disorders: challenges in the later phases of the recovery process: a qualitative study of patients’ experiences. Scand J Caring Sci. 2013;27(1):92–8.

Charmaz K. The power of constructivist grounded theory for critical inquiry. Res Methods Eval. 2016;23(1):34–45.

Williams K, King J, Fox JR. Sense of self and anorexia nervosa: a grounded theory. Psychol Psychother. 2016;89(2):211–28.

Smith JA. Reflecting on the development of interpretive phenomenological analysis and its contribution to qualitative research in psychology. Qual Res Psychol. 2008;1(1):39–54.

Fox AP, Larkin M, Leung N. The personal meaning of eating disorder symptoms: an interpretative phenomenological analysis. J Health Psychol. 2011;16(1):116–25.

Henry D, Dymnicki AB, Mohatt N, Allen J, Kelly JG. Clustering methods with qualitative data: a mixed-methods approach for prevention research with small samples. Prev Sci. 2015;16(7):1007–16.

Saunders JF, Eaton AA. Social comparisons in eating disorder recovery: using photovoice to capture the sociocultural influences on women’s recovery. Int J Eat Disord. 2018;51(12):1361–6.

Saunders JF, Eaton AA, Aguilar S. From self(ie)-objectification to self-empowerment: the meaning of selfies on social media in eating disorder recovery. Comput Human Beh. 2020;111:106420.

Wang C, Burris MA. Photovoice: concept, methodology, and use for participatory needs assessment. Health Educ Behav. 1997;24(3):369–87.

LaMarre A, Rice C. Embodying critical and corporeal methodology: digital storytelling With young women in eating disorder recovery. Forum Qual Soc Res. 2016;17:2.

Bartel H. A ‘Girls’ Illness?’ Using narratives of eating disorders in men and boys in healthcare education and research. In: Arts based health care research: a multidisciplinary perspective. Cham: Springer; 2022. p. 69-84.

Proszynski J. Interior designs: demolition and reconstruction of the self in eating disorder recovery through transformable collage book making (Doctoral dissertation, Pratt Institute). 2022.

Marzola E, Abbate-Daga G, Gramaglia C, Amianto F, Fassino S, Mobini S (ReviewingEditor). A qualitative investigation into anorexia nervosa: The inner perspective. Cogent Psychol 2015;2:1. https://doi.org/10.1080/23311908.2015.1032493 .

Stewart C, Konstantellou A, Kassamali F, McLaughlin N, Cutinha D, Bryant-Waugh R, et al. Is this the ‘new normal’? A mixed method investigation of young person, parent and clinician experience of online eating disorder treatment during the COVID-19 pandemic. J Eat Disord. 2021;9(1):78.

Leehr EJ, Schag K, Dresler T, Grosse-Wentrup M, Hautzinger M, Fallgatter AJ, et al. Food specific inhibitory control under negative mood in binge-eating disorder: Evidence from a multimethod approach. Int J Eat Disord. 2018;51(2):112–23.

Bartholome LT, Raymond NC, Lee SS, Peterson CB, Warren CS. Detailed analysis of binges in obese women with binge eating disorder: Comparisons using multiple methods of data collection. Int J Eat Disord. 2006;39(8):685–93.

Mehak A, Racine SE. Understanding “feeling fat” and its underlying mechanisms: the importance of multimethod measurement. Int J Eat Disord. 2020;53(9):1400–4.

Harrop E. “Maybe i really am too fat to have an eating disorder”: A mixed methods study of weight stigma and healthcare experiences in a diverse sample of patients with atypical anorexia (Doctoral dissertation). Seattle: University of Washington; 2020.

Fusar-Poli P, Rutigliano G, Stahl D, Davies C, Bonoldi I, Reilly T, et al. Development and validation of a clinically based risk calculator for the transdiagnostic prediction of psychosis. JAMA Psychiatry. 2017;74(5):493–500.

Nobel L, Mayo NE, Hanley J, Nadeau L, Daskalopoulou SS. MyRisk_stroke calculator: a personalized stroke risk assessment tool for the general population. J Clin Neurol. 2014;10(1):1–9.

Gail MH, Brinton LA, Byar DP, Corle DK, Green SB, Schairer C, et al. Projecting individualized probabilities of developing breast cancer for white females who are being examined annually. J Natl Cancer Inst. 1989;81(24):1879–86.

Perlis RH. Pharmacogenomic testing and personalized treatment of depression. Clin Chem. 2014;60(1):53–9.

Hafeman DM, Merranko J, Goldstein TR, Axelson D, Goldstein BI, Monk K, et al. Assessment of a person-level risk calculator to predict new-onset bipolar spectrum disorder in youth at familial risk. JAMA Psychiatry. 2017;74(8):841–7.

Birmaher B, Merranko JA, Goldstein TR, Gill MK, Goldstein BI, Hower H, et al. A risk calculator to predict the individual risk of conversion from subthreshold bipolar symptoms to bipolar disorder I or II in youth. J Am Acad Child Adolesc Psychiatry. 2018;57(10):755–63 e4.

Birmaher B, Merranko JA, Gill MK, Hafeman D, Goldstein T, Goldstein B, et al. Predicting personalized risk of mood recurrences in youths and young adults with bipolar spectrum disorder. J Am Acad Child Adolesc Psychiatry. 2020;59(10):1156–64.

Goldstein TR, Merranko J, Hafeman D, Gill MK, Liao F, Sewall C, et al. A risk calculator to predict suicide attempts among individuals with early-onset bipolar disorder. Bipolar Disord. 2022.

Fuglset TS, Landro NI, Reas DL, Ro O. Functional brain alterations in anorexia nervosa: a scoping review. J Eat Disord. 2016;4:32.

Phillipou A, Beilharz F. Should we shed the weight criterion for anorexia nervosa? Aust N Z J Psychiatry. 2019;53(6):501–2.

Kenny TE, Lewis SP. Reconceptualizing recovery: integrating lived experience perspectives into traditional eating disorder recovery frameworks. Psychiatr Serv. 2021;72(8):966–8.

Ramjan LM, Hay P, Fogarty S. Benefits of a mentoring support program for individuals with an eating disorder: a proof of concept pilot program. BMC Res Notes. 2017;10(1):709.

Wolf M, Thies F, Kordy H. Language use in eating disorder blogs: psychological implications of social online activity. J Lang Soc Psychol. 2013;32(2):212–26.

Download references

Acknowledgements

The authors wish to thank our colleagues in ED recovery.

None received.

Author information

Authors and affiliations.

Department of Psychiatry, Eating Disorders Center for Treatment and Research, University of California at San Diego School of Medicine, 4510 Executive Drive, San Diego, CA, 92121, USA

Heather Hower

Department of Health Services, Policy, and Practice, Hassenfeld Child Innovation Institute, Brown University School of Public Health, 121 South Main Street, Providence, RI, 02903, USA

School of Psychology, Massey University, North Shore, Private Bag 102-904, Auckland, 0632, New Zealand

Andrea LaMarre

Clinical Psychology Graduate Program, Ruppin Academic Center, 4025000, Emek-Hefer, Israel

Rachel Bachner-Melman

School of Social Work, Hebrew University of Jerusalem, Mt. Scopus, 9190501, Jerusalem, Israel

Graduate School of Social Work, University of Denver, 2148 S High Street, Denver, CO, 80208, USA

Erin N. Harrop

University of Kansas School of Medicine, 1010 N Kansas St, Wichita, KS, 67214, USA

Beth McGilley

Department of Psychology, Clinical Child and Adolescent Psychology, University of Guelph, 50 Stone Road East, Guelph, ON, N1G 2W1, Canada

Therese E. Kenny

You can also search for this author in PubMed   Google Scholar

Contributions

HH contributed to the conceptualization of the manuscript, the writing and editing of the abstract, introduction, quantitative text and table, mixed-methods, creative methods, multi-methods, future research, and conclusions sections, EndNote reference library/citations, formatting of the manuscript, and overall final revisions. AL contributed to the conceptualization of the manuscript, the writing and editing of the abstract, introduction, qualitative text and table, mixed-methods, creative methods, and conclusion sections, and overall final revisions. RBM contributed to the conceptualization of the manuscript, the writing and editing of the abstract, introduction, quantitative, and qualitative sections, and overall final revisions. EH contributed to the conceptualization of the manuscript, the writing and editing of the abstract, introduction, qualitative text and table sections, and overall final revisions. BM contributed to the conceptualization of the manuscript, the writing and editing of the abstract, introduction, qualitative, and mixed-methods sections, and overall final revisions. TK contributed to the conceptualization of the manuscript, the writing and editing of the epistemology and ontology sections, and discussion of these concepts through other sections, as well as overall final revisions. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Heather Hower .

Ethics declarations

Ethics approval and consent to participate, consent for publication, competing interests.

RBM is a Senior Editor of this Journal. All other authors declare that they have no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Hower, H., LaMarre, A., Bachner-Melman, R. et al. Conceptualizing eating disorder recovery research: Current perspectives and future research directions. J Eat Disord 10 , 165 (2022). https://doi.org/10.1186/s40337-022-00678-8

Download citation

Received : 05 September 2022

Accepted : 25 October 2022

Published : 15 November 2022

DOI : https://doi.org/10.1186/s40337-022-00678-8

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Quantitative research
  • Qualitative research
  • Multi-methods research

Journal of Eating Disorders

ISSN: 2050-2974

recovery model research paper

recovery model research paper

Academia.edu no longer supports Internet Explorer.

To browse Academia.edu and the wider internet faster and more securely, please take a few seconds to  upgrade your browser .

  •  We're Hiring!
  •  Help Center

Recovery Model of Mental Health

  • Most Cited Papers
  • Most Downloaded Papers
  • Newest Papers
  • Save to Library
  • Recovery in Mental Health Follow Following
  • Peer Support in Recovery of People with Mental Health Challenges Follow Following
  • Recovery model in mental illness Follow Following
  • Journal of Psychiatric and Mental Health Nursing Follow Following
  • Family Strengths and Resilience Follow Following
  • Recovery in Mental Illness Follow Following
  • Phenomenology- Mind/Body Problems/ Merleau-Ponty's Philosophical Thought/Phenomenology and Embodiment Follow Following
  • Institutionalization Theory Follow Following
  • Psychosocial Interventions and Mental Health Nursing Follow Following
  • Greg madison Follow Following

Enter the email address you signed up with and we'll email you a reset link.

  • Academia.edu Publishing
  •   We're Hiring!
  •   Help Center
  • Find new research papers in:
  • Health Sciences
  • Earth Sciences
  • Cognitive Science
  • Mathematics
  • Computer Science
  • Academia ©2024

This paper is in the following e-collection/theme issue:

Published on 16.4.2024 in Vol 26 (2024)

This is a member publication of University of Cambridge (Jisc)

Factors Influencing Recovery From Pediatric Stroke Based on Discussions From a UK-Based Online Stroke Community: Qualitative Thematic Study

Authors of this article:

Author Orcid Image

Original Paper

  • Charlotte Howdle 1   ; 
  • William James Alexander Wright 1 , BChir, MB, MA   ; 
  • Jonathan Mant 2 , MBBS, MA, MSc, MD   ; 
  • Anna De Simoni 2, 3 , MBBS, PhD, MRCGP  

1 School of Clinical Medicine, University of Cambridge, Cambridge, United Kingdom

2 Department of Public Health and Primary Care, University of Cambridge, Cambridge, United Kingdom

3 Wolfson Institute of Population Health, Queen Mary University of London, London, United Kingdom

Corresponding Author:

Anna De Simoni, MBBS, PhD, MRCGP

Wolfson Institute of Population Health

Queen Mary University of London

58 Turner Street

Centre for Primary Care

London, E1 2AB

United Kingdom

Phone: 44 207882 ext 252

Email: [email protected]

Background: The incidence of stroke in children is low, and pediatric stroke rehabilitation services are less developed than adult ones. Survivors of pediatric stroke have a long poststroke life expectancy and therefore have the potential to experience impairments from their stroke for many years. However, there are relatively few studies characterizing these impairments and what factors facilitate or counteract recovery.

Objective: This study aims to characterize the main barriers to and facilitators of recovery from pediatric stroke. A secondary aim was to explore whether these factors last into adulthood, whether they change, or if new factors impacting recovery emerge in adulthood.

Methods: We performed a qualitative thematic analysis based on posts from a population of participants from a UK-based online stroke community, active between 2004 and 2011. The analysis focused on users who talked about their experiences with pediatric stroke, as identified by a previous study. The posts were read by 3 authors, and factors influencing recovery from pediatric stroke were mapped into 4 areas: medical, physical, emotional, and social. Factors influencing recovery were divided into short-term and long-term factors.

Results: There were 425 posts relating to 52 survivors of pediatric stroke. Some survivors of stroke posted for themselves, while others were talked about by a third party (mostly parents; 31/35, 89% mothers). In total, 79% (41/52) of survivors of stroke were aged ≤18 years and 21% (11/52) were aged >18 years at the time of posting. Medical factors included comorbidities as a barrier to recovery. Medical interventions, such as speech and language therapy and physiotherapy, were also deemed useful. Exercise, particularly swimming, was deemed a facilitator. Among physical factors, fatigue and chronic pain could persist decades after a stroke, with both reported as a barrier to feeling fully recovered. Tiredness could worsen existing stroke-related impairments. Other long-standing impairments were memory loss, confusion, and dizziness. Among emotional factors, fear and uncertainty were short-term barriers, while positivity was a major facilitator in both short- and long-term recovery. Anxiety, grief, and behavioral problems hindered recovery. The social barriers were loneliness, exclusion, and hidden disabilities not being acknowledged by third parties. A good support network and third-party support facilitated recovery. Educational services were important in reintegrating survivors into society. Participants reported that worrying about losing financial support, such as disability allowances, and difficulties in obtaining travel insurance and driving licenses impacted recovery.

Conclusions: The lived experience of survivors of pediatric stroke includes long-term hidden disabilities and barriers to rehabilitation. These are present in different settings, such as health care, schools, workplaces, and driving centers. Greater awareness of these issues by relevant professional groups may help ameliorate them.

Introduction

A pediatric stroke is classified as either perinatal (occurring when the child is 20 weeks in gestation to 28 days after birth) or childhood (from 29 days after birth to 18 years). There are >400 cases of pediatric stroke in the United Kingdom annually [ 1 ]. Pediatric stroke can be a debilitating disease, leaving survivors and their families coping with persisting issues during their recovery. These problems evolve and become more numerous as the survivor grows up, with as many as 75% of families of a child with stroke having at least 1 unmet need [ 2 ].

There is limited literature on factors that impact recovery. Survivors of pediatric stroke often develop comorbidities. Studies have found that survivors have an increased incidence of the following: epilepsy [ 3 - 5 ], attention-deficit/hyperactivity disorder [ 3 , 5 ], depression [ 5 ], anxiety [ 4 , 5 ], and autism [ 6 , 7 ], compared with their nonstroke counterparts. Fatigue [ 8 ] and behavioral problems [ 4 , 9 ] also hinder recovery. Mental health support [ 10 ] and follow-up neuropsychological assessments [ 10 , 11 ] positively influence outcomes. Qualitative studies with parents revealed that many disabilities experienced by stroke survivors are hidden [ 9 ], resulting in them not receiving the support they need. 

There are also emotional barriers that impact recovery. Pediatric stroke negatively affects the family of a survivor, with reports of poor parental mental health [ 4 , 11 ], guilt [ 4 , 12 , 13 ], and uncertainty about the future [ 12 ]. Having good parental welfare is important—parents are the primary caregivers following their child’s stroke, and their well-being impacts their child’s recovery. Good family functioning [ 4 ] also facilitates recovery from childhood stroke. Parents reported that treating the child as “normal” aided recovery [ 14 ]. A supportive social network, from close family and friends [ 14 ] to the wider society [ 13 ], is important for both parents and children to facilitate recovery.

Certain social factors are known to impact recovery. Qualitative reviews asking about patients’ experiences of health care showed that after discharge, parents felt abandoned by professionals [ 9 , 12 ]. Health care services may not be flexible, they may be difficult to access, and parents may not know where they can access therapy for their child [ 12 ]. When health care staff could be approached, parents may not have known what questions to ask [ 15 ]. Clear communication with parents by medical professionals about the causes of the stroke and events occurring around the stroke has been identified as an important issue [ 4 ]. Patients have appreciated positivity from clinicians [ 14 ], close and ongoing medical support [ 14 , 15 ], involvement with goal-setting approaches [ 16 , 17 ], and continuity of care. There is limited medical awareness or literature to support parents during their child’s recovery from stroke [ 2 , 9 , 12 ]. Support from charities and both in-person and online community groups partly addresses this issue [ 4 , 13 , 15 ]. Accessing disability aid also appears to facilitate recovery from childhood stroke [ 15 ].

A growing body of literature supports the potential of online health communities to provide opportunities for individuals to share their personal experiences and learn from others with similar conditions [ 18 , 19 ]. These communities also serve as a reliable and novel source of information about patients’ unmet needs [ 20 , 21 ], complementing more traditional research methods [ 22 , 23 ].

There are an increasing number of studies that explore the barriers to and facilitators of recovery from pediatric stroke, considering that these factors may be numerous and long lasting as children have a long poststroke life expectancy. However, there are relatively few studies that characterize the long-term impact of stroke on children; currently, the longest time frame studied is 5 years after stroke [ 24 ]. In May 2017, the Royal College of Paediatrics and Child Health in the United Kingdom published The Stroke in Childhood Clinical Guidelines , which had a set of “research recommendations” that included “reviewing the complications that children experience” following pediatric stroke to assess the “ rehabilitation needs of pediatric stroke patients” and evaluate the “long-term outcomes” of survivors of pediatric stroke [ 25 ]. This study aims to address these recommendations by characterizing what survivors of pediatric stroke report as the main barriers and facilitators to their recovery in a UK-based online stroke community. As a secondary aim, the study also explores whether these factors last into adulthood or not, whether they change, or if new factors impacting recovery emerge in adulthood.

Study Design

We conducted a qualitative thematic analysis on posts from a pediatric stroke population within a UK-based online stroke community.

The analysis used the archived TalkStroke online community, a UK-based, moderated online community hosted on the Stroke Association website from 2004 to 2011. In total, the TalkStroke archive contains 22,173 posts written by 2583 unique usernames. A previous study by De Simoni et al [ 18 ] identified 58 usernames that posted about experiences of pediatric stroke, contributing to a total of 469 posts. We excluded some users: 2 after further analysis revealed that their age at stroke was >18 years and a further 4 users because their age at the time of posting was unknown. A sample of 52 users remained. The characteristics of the survivors of pediatric stroke, including demographics, employment, education, stroke type, initial impairments as well as impairments at the time of posting, support needs, and independence, were retrieved from the data set of a previous study [ 18 ].

Data Analysis

A deductive approach was used to develop predetermined themes before the start of the analysis. Factors (barriers or facilitators) were split into 4 themes based on recommendations by the Clinical Guidelines for Stroke in Childhood [ 25 ]: medical, physical, emotional, and social.

All relevant posts were collected in an Excel (Microsoft Inc) spreadsheet. WJAW and CH read all the posts to become familiar with the data. Themes were further split into subcategories using a data-driven approach by applying thematic analysis, as described by Braun and Clarke [ 26 ]. Coding was discussed until agreement was reached and a final coding framework was agreed upon. Each individual post was considered on its own, outside the context of the thread to which the posts belonged.

To consider the potentially different perspectives of participants over time, the posts were split into 2 categories: whether the survivor of pediatric stroke was aged ≤18 years or >18 years at the time of posting. This provided an assessment of whether the factors impacting recovery were short-term or long-term factors. We defined “long-term factors” as those affecting participants into adulthood and “short-term factors” as those impacting patients from the time immediately following a stroke until they reach 18 years of age. Sometimes, third parties (ie, parents or a member of the wider family) wrote on behalf of the survivor of pediatric stroke.

In this analysis, the term recovery describes the improvement of any aspect of stroke-related impairment. Rehabilitation is defined in concordance with the definition of the World Health Organization: “interventions designed to optimize functioning and reduce disability in individuals with health conditions in interaction with their environment” [ 27 ].

Patient and Public Involvement

A 22-year-old survivor of pediatric stroke (when aged 0 years) was contacted after the analysis was completed. The patient and public involvement (PPI) participant was an acquaintance of a coauthor’s friend and was approached informally. We obtained written consent for her contribution to the work to be included in the paper.

The initial results were written and sent to the PPI participant, who read and provided insightful comments that highlighted important subcategories and informed our Discussion section.

Ethical Considerations

The Stroke Association collected the data from the archived TalkStroke forum and provided the data set, together with their permission, for analysis for research purposes. The data are stored through the University of Cambridge Clinical School’s Secure Data Hosting Service, with reference S0126—Stroke Needs & Exp. This analysis was assessed by the University of Cambridge institutional review board and exempted from ethics approval, provided that permission to use the data was granted by the Stroke Association. At the time of registering with the community, users agreed that their data would be public. However, to protect the identity and intellectual property of participants, this analysis does not use direct quotes; instead, quotes are paraphrased. A more detailed account of the ethics involved with research on the TalkStroke archives is described in the study by De Simoni et al [ 18 ].

Participants’ Characteristics

In total, 79% (41/52) of users were aged ≤18 years at the time of participation, contributing a total of 273 posts; 21% (11/52) of users were aged >18 years and contributed 152 posts. The majority of data from the ≤18-year-old group were collected through third-party users (35/41, 85%). Of these, most were written by the mother of the survivor (31/35, 89%). Data from the >18-year-old group were reported firsthand by adult survivors of pediatric stroke ( Table 1 ).

Among the 11 participants who were aged >18 years at the time of posting, 2 (18%) held a driving license, 1 (9%) was considering applying for one, and 1 (9%) stated they do not drive. In total, 4 (36%) participants were university graduates or attending university, 1 (9%) was in part-time employment, and 1 (9%) was in full-time employment.

a Age at time of participation in the online stroke community.

b x: time or age between the 2 values.

c N/A: not applicable.

Causes of Stroke and Resulting Impairments

Of the 52 survivors, 13 (25%) were survivors of right-sided strokes, 22 (42%) were survivors of left-sided strokes, 2 (4%) were survivors of stroke affecting both sides, and 15 (29%) were not reported. The causes of stroke were reported for a small number of participants: 4 (8%) were after surgery (3/4, 75% was due to cardiac operations and 1/4, 25% was unknown); 6 (11%) were after infection (1/6, 17% was meningitis; 3/6, 50% was chicken pox; 1/6, 17% was herpes encephalopathy; and 1/6, 17% was maternal shingles during gestation); 2 (4%) were ischemic; 3 (6%) were hemorrhagic; 1 (2%) was linked to acute lymphoblastic leukemia treatment; 2 (4%) were dissections; 1 (2%) was linked to a brain tumor; 1 (2%) was linked to arterial stenosis; 1 (2%) was linked to a septal defect; 1 (2%) was linked to an oral contraceptive pill; and 30 (58%) were not reported.

The initial impairments after stroke included the following: hemiplegia, hemiparesis, poststroke epilepsy, visual disturbances, tiredness, increased emotions, dystonia, dysarthria, facial droop, memory impairments, headaches, and no impairments. The difference between initial impairments and residual impairments at the time of posting varied greatly for users. The time between stroke and participation in the community ranged from 2 weeks to 46 years.

The subcategories of the 4 themes are presented in Table 2 , which shows whether the factors are short term, mostly reported by parents discussing their children with stroke, or long term, as reported by adult survivors of pediatric strokes.

Medical Factors

Comorbidities.

Epilepsy and depression were most commonly mentioned. Parents found it tough to cope with the additional diagnosis of epilepsy alongside pediatric stroke, with some reporting seeing their children getting more ill rather than better:

A parent writes they were completely shattered as their child had already suffered so much. [Mother; participant 30; ≤18-year-old group; stroke at the age of 11 years; 2 years since the stroke]
One survivor wrote their depression sets them in a really low state of mind, where they cannot control their emotion. [Survivor; participant 47; ≤18-year-old group; stroke at the age of 15 years; 1 year since the stroke]
Another, 35 years after their stroke, queried if it was possible to still feel depressed due to the stroke. [Survivor; participant 36; >18-year-old group; stroke at the age of 0 years; 35 years after the stroke]

Medical Interventions

The survivors credited physiotherapy and speech therapy with helping them regain functionality and speech, respectively. The only downside mentioned was long waiting times to access services:

One parent advised others to get on top of physio immediately, saying that she went privately as waiting lists were long. [Mother; participant 12; ≤18-year-old group; stroke at the age of 1 year; 11 years after the stroke]
One adult survivor writes that having physiotherapy sessions helped her walk. [Survivor; participant 36; >18-year-old group; stroke at the age of 0 years; 35 years after the stroke]

Exercise was an activity recommended by medical practitioners that aided recovery. Parents and survivors mentioned the importance of repeatedly doing both general exercise and muscle-specific exercises to prevent muscle wasting, build muscle strength, help induce sleep, and increase fitness. Comments from the >18-year-old group also supported the utility of exercise, stressing the importance of regular daily activity. Swimming was written about consistently; it was recommended by parents as a way to get their children’s limbs to move, and it was noted that gains in swimming ability became an indicator of recovery and a source of excitement for the survivors and their families:

One individual advised parents to make a child do their exercises every day. [Survivor; participant 52; >18-year-old group; stroke at the age of 17 years; 21 years after the stroke]
One parent wrote that swimming is a great way of exercising and moving their child’s limbs without trauma. [Mother; participant 11; ≤18-year-old group; stroke at the age of 1 year; 1 year after the stroke]
Another wrote their child had 1 to 1 swimming lessons and could nearly do breaststroke again. [Mother; participant 25; ≤18-year-old group; stroke at the age of 8 years; 0 years after the stroke]

Other medical interventions that the participants mentioned were the use of Botox for tightness in muscles. Although it was often useful, it did not always help. In addition, the participants aged >18 years endorsed the use of SaeboFlex [ 28 ] for regaining functionality, tai chi for mental health, and quinine for muscle cramps and spasms.

Physical Factors

Fatigue was the most commonly reported physical barrier to recovery, with some parents writing that it was their main concern. When sleep duration was reported, it was at least 12 hours each night. The users told each other that tiredness was often caused by the large effort that the survivors were putting into their recovery. Fatigue was also a long-term factor:

A parent advised another that tiredness as a result of stroke is normal. [Mother; participant 26; ≤18-year-old group; stroke at the age of 9 years; 0 years after the stroke]
One user reported that there was underlying tiredness throughout middle age, when they were working full time. In another post, they wrote that having a stroke resulted in less stamina, tolerance and energy in the context of noisy, busy backgrounds. [Survivor; participant 45; >18-year-old group; stroke at the age of 15 years; 46 years after the stroke]

When a survivor became tired, this caused a worsening in disability. This affected speech, movement, and coordination:

A parent reported that when their child became tired on holiday, they were sad to see that their leg and arms were dragging. [Mother; participant 12; ≤18-year-old group; stroke at the age of 1 year; 0 years after the stroke]

Pain was another commonly cited barrier preventing people from feeling fully recovered. Sites of pain were varied, including back pain, headaches, and limb muscle cramps. Pain emerged as a long-term consequence of a pediatric stroke. When the users discussed their pain with medical professionals, they were told that it was a result of their stroke and that it would ease with time. Being cold exacerbated chronic pain both in the short and long term:

One user reported that their headaches were seriously affecting their everyday life. [Survivor; participant 47; ≤18-year-old group; stroke at the age of 15 years; 1 year after the stroke]
A user still has pain in their joints 35 years poststroke. [Survivor; participant 36; >18-year-old group; stroke at the age of 0 years; 35 years after the stroke]
A parent wrote that their child’s affected limbs were more painful in the cold weather. [Mother; participant 32; >18-year-old group; stroke at the age of 13 years; 1 year after the stroke]

Neurological Sequelae

Poststroke memory loss, confusion, and dizziness were the barriers to recovery both in the short and long term:

A parent described her child’s memory loss as causing as much trouble as the arm or foot. [Mother; participant 32; ≤18-year-old group; stroke at the age of 13 years; 1 year after the stroke]
An individual 3 years poststroke commented that his eyes are dizzy. [Survivor; participant 41; >18-year-old group; stroke at the age of 11 years; 30 years after the stroke]

Emotional Factors

Adult survivors of pediatric stroke stressed that positivity was an important factor in ensuring a successful recovery. It was mentioned less often by parents and younger survivors of pediatric stroke. However, humor was brought up by both groups as an important facilitator in coping with the consequences of stroke. Many possible fun activities were exchanged on the site. There was a large proportion of music-related activities, for example, singing songs with hand puppets:

One survivor stresses staying positive is the key to a successful recovery. [Survivor; participant 49; >18-year-old group; stroke at the age of 17 years; 2 years after the stroke]
One parent writes their family tries to laugh themselves through bad times, rather than cry again. [Mother; participant 32; ≤18-year-old group; stroke at the age of 13 years; 1 year after the stroke]
One parent wrote that music is a great mental stimulator as well as being fun. [Mother; participant 12; ≤18-year-old group; stroke at the age of 1 year; 0 years after the stroke]

Grief and Bereavement

Following a pediatric stroke, the survivors wrote that they viewed themselves as different people and grieved for the person they once were. The families of survivors also commented that they experienced similar emotions and that the stroke had a long-term effect on the whole family:

One parent commented they have a different child now. [Mother; participant 31; ≤18-year-old group; stroke at the age of 11 years; 5 years after the stroke]
Another survivor queried whether anyone on the forum felt that stroke was like a loss in the family and caused a grieving process. [Survivor; participant 51; >18-year-old group; stroke at the age of 17 years; 4 years after the stroke]
One parent commented that they don’t think any of their family will ever be the same as before the stroke. [Mother; participant 32; ≤18-year-old group; stroke at the age of 13 years; 1 year after the stroke]

Fear of uncertainty and the unknown was commonly cited as a barrier to recovery. Parents wrote that the fear of another stroke event, how their child will grow up and fit into society, and having no explanation of the cause of the stroke were particularly difficult issues. In contrast, adult survivors encouraged users to accept the stroke and not let fear stand in the way of recovery:

A parent wrote that it was hard to comprehend the unknown future. [Mother; participant 11; ≤18-year-old group; stroke at the age of 1 year; 1 year after the stroke]
One adult survivor wrote that acceptance of stroke is the first stage of the healing process and survivors must move on and get on with their lives. [Survivor; participant 46; >18-year-old group; stroke at the age of 15 years; 28 years after the stroke]

Both parents and survivors described the scenarios that caused upset by reminding them of the trauma surrounding the stroke event:

One parent writes that her daughter was admitted to hospital and it brings back too many memories which makes their calm slip a little inside. [Mother, participant 12; ≤18-year-old group; stroke at the age of 1 year; 0 years after the stroke]
An adult survivor training to be a health care professional writes that having a placement on a stroke unit is difficult to cope with. [Survivor; participant 51; >18-year-old group; stroke at the age of 17 years; 4 years after the stroke]

Behavioral Problems

Behavioral problems were commonly mentioned. Parents reported mood swings with negative emotions, such as getting upset, anger, and frustration. There were multiple posts trying to rationalize why these changes occur, possible reasons being parental spoiling of the child either before or after the stroke and typical teenage mood swings. Adult survivors of pediatric stroke mentioned some lasting behavioral problems:

A parent wrote that their child gets upset with how she feels and that her brain seems to tell her things that she can’t cope with. [Mother; participant 30; ≤18-year-old group; stroke at the age of 11 years; 2 years after the stroke]
One survivor wrote that he takes himself to bed when he gets really grumpy. [Survivor; participant 52; >18-year-old group; stroke at the age of 17 years; 21 years after the stroke]
Another survivor queries whether it is possible to have mood swings and feel low 35 years after a stroke as she is struggling. [Survivor; participant 36; >18-year-old group; stroke at the age of 0 years; 35 years after the stroke]

Social Factors

Support network.

Parents stressed the importance of supporting their children throughout the stroke recovery process. Support was appreciated by the survivors of stroke, who reiterated the importance of love and support in motivating them during recovery:

One parent wrote it is completely down to the family to get their child through recovery. [Mother; participant 12; ≤18-year-old group; stroke at the age of 1 year; 0 years since the stroke]

An adult survivor of pediatric stroke wrote that they fought every inch of the way with the love and support of their parents. [Survivor; participant 39; >18-year-old group; stroke at the age of 8 years; 23 years since the stroke]
A user mentioned their friend by name and thanked them for getting them over the mental side of things. [Survivor; participant 42; >18-year-old group; stroke at the age of 13 years; 7 years after the stroke]

Loneliness emerged in many contexts as both a short-term and a long-term barrier to recovery from stroke. Loneliness made the survivors feel different from those around them, which negatively impacted their well-being. The survivors felt isolated from fellow survivors of stroke, friends, family, and peers:

A male survivor wrote he struggled to find any help and support with his rehabilitation. [Survivor; participant 43; >18-year-old group; stroke at the age of 13 years; 7 years after the stroke]
A female survivor commented she never talks about her stroke to her friends because she doesn’t want to be judged. [Survivor; participant 36; >18-year-old group; stroke at the age of 0 years; 35 years after the stroke]
The parent of a survivor writes that her child’s class look at him differently. [Mother; participant 20; ≤18-year-old group; stroke at the age of 3 years; 2 years after the stroke]

Loneliness had long-lasting effects on survivors of stroke:

One member of the >18 group reported they felt alone most of their life. They then go on to tell a user they have done the best thing joining the site. [Survivor; participant 36; >18-year-old group; stroke at the age of 0 years; 35 years after the stroke]

Exclusion or Bullying

There were reports of bullying and exclusion behaviors among children at school. Furthermore, parents reported that social restrictions, such as not allowing the survivors to ride a bicycle or go out with friends, were a barrier to recovery:

A survivor’s mother wrote that her daughter is taunted at school by her own friends for having a stroke. [Mother; participant 26; ≤18-year-old group; stroke at the age of 9 years; 0 years after the stroke]
One parent wrote that the restrictions placed on her son prompted him to tell them that he may as well be dead. [Mother; participant 22; ≤18-year-old group; stroke at the age of 5 years; 4 years after the stroke]

Hidden Disabilities

Being doubted about their degree of disability was reported by several adult survivors of pediatric stroke. They described the medical staff and the public as not understanding the hidden effects of stroke and assuming that if there are no visible impairments, they are recovered. This discredited the survivor’s struggle for recovery and made them feel upset:

A survivor wrote how she told a midwife about her stroke and the midwife’s response was that because they hadn’t noticed at first glance, it couldn’t have affected the survivor that badly. The survivor commented how angry that made her feel. [Survivor; participant 36; >18-year-old group; stroke at the age of 0 years; 35 years after the stroke]
A survivor reported people discrediting their tiredness, accusing them of being selfish or lazy. [Survivor; participant 51; >18-year-old group; stroke at the age of 17 years; 4 years after the stroke]

There were mixed reviews about the education services accessed; however, interaction with the education sector was highlighted as important in getting the survivor of pediatric stroke back into school and therefore integrating back into society:

One parent writes that her child’s new school is not dealing with her needs very well, and they have had to go in on several occasions. [Mother; participant 32; >18-year-old group; stroke at the age of 13 years; 1 year after the stroke]

There was concern from parents of survivors of stroke as to whether their children could drive. There was much reassurance and advice given by other parents and survivors in the >18-year-old group:

One parent writes that her child had his stroke at 11 and was then 16 and wondered if he would be able to drive. [Mother; participant 31; ≤18-year-old group; stroke at the age of 11 years; 5 years after the stroke]
One survivor commented he overcame medical predictions and had been driving since the age of 21. [Survivor; participant 46; >18-year-old group; stroke at the age of 15 years; 28 years after the stroke]

Parents reported that obtaining travel insurance for their child was difficult. Many companies did not provide insurance at all; they only insured per trip and not for longer periods as parents wanted, and the price of insurance was high, deterring the families from traveling:

One parent commented that they found it mind-blowing that they must pay so much and no one else will cover them. [Survivor; participant 52; >18-year-old group; stroke at the age of 17 years; 21 years after the stroke]

Third-Party Support

The members often recommended external resources to each other. These were most often helplines or information on websites. Organizations that were mentioned were the Stroke Association [ 1 ], HemiHelp [ 29 ], Maypole project [ 30 ], Different Strokes [ 31 ], Mobilise [ 32 ], Disability Information and Advice Line [ 33 ], and Headway [ 34 ].

Financial Support

Financial support for disability was available for survivors of pediatric stroke and was identified as helpful. However, there was discussion among adult survivors of pediatric stroke who expressed fear of having financial support taken away from them:

One survivor wrote that they had read an article about how people with mental health issues are more likely to fail the assessment test, as they have multiple symptoms which vary and she was worried that this would disadvantage stroke survivors also. [Survivor; participant 45; >18-year-old group; stroke at the age of 15 years; 46 years after the stroke]

PPI Feedback

The analysis was read by a survivor of multiple childhood strokes while aged <1 year who is now studying at a university. She reported strongly agreeing with uncertainty about the future, the benefits of swimming, the hidden disabilities because of stroke, and the lack of public awareness of pediatric stroke. Some quotes are reported in the subsequent section to illustrate the PPI feedback:

  • Regarding swimming:
I was also told it was important for me to swim to help prevent any damage as I developed so swam from 6 months on.
  • Uncertainty about the future:
This strongly resonates with me and my family. For my parents it was a huge unknown as the doctors couldn’t tell them if I would be able to walk, speech or do well in school. They also couldn’t tell if I was going to have another one, but if I didn’t have another by 12 I should be clear from not having another.
  • Hidden disabilities:
I would agree hidden effects of a stroke are not spoken about. In the main all I was warned of was physical disabilities and those which would be major hindrances in my life. By this I mean walking, talking, learning and doing physical exercise/sport well. Once it was clear at ~7 or 8 year old I was able to conduct those tasks no further check ups were taken and no-one checked for any other less major/obvious effects. As for the public, I would strongly agree many have no idea of the effects of strokes, even I didn’t realise tiredness was an effect until I read this paper! I have an absolutely terrible memory and I do think if this is very likely down to my stroke and this I would say is a hidden disability. I’m almost certain my stroke affected the memory side of my brain so it would make some sense. However, I assume few members of the public would understand this and no medical staff checked or mentioned smaller effects other than physical movement/speech/learning.

Principal Findings

Medical, physical, emotional, and social themes were identified as impacting the recovery from pediatric stroke. Exercise, swimming, speech therapy, and physiotherapy were the medical factors that facilitated recovery, whereas having a comorbidity hindered recovery. A novel finding of this study is that “hidden” physical impairments, such as fatigue, pain, memory loss, confusion, and dizziness, affected survivors of pediatric stroke in the long term, with the lack of awareness of these impairments by the general public and by professionals in health care, school, and workplace settings also hindering recovery. Grief, fear of the future, restrictions on the child’s life, and behavioral problems were the emotional factors that slowed recovery, whereas positivity and a good support network aided recovery. Isolation, hidden disabilities, triggering events, finding travel insurance, and fear of not being able to drive were the social factors hindering recovery. Third-party support and financial aid were facilitators of recovery, albeit there was a fear of losing financial aid as survivors of pediatric stroke aged.

Identifying the factors that impact recovery from pediatric stroke is important, as there is the potential for survivors to live for many decades after the stroke, and therefore, they may live many years with disabling factors. Our study found that fatigue persists many decades after stroke, which expands on current literature that fatigue is an identified barrier up to 5 years after the stroke [ 8 ]. In addition, tiredness was reported to exacerbate stroke-related disabilities. This relationship has been reported in adult brain injury [ 35 ], but to the best of our knowledge, this is the first time it has been reported in pediatric stroke. In addition, this study is the first to report chronic pain as a barrier to recovery in pediatric stroke, despite chronic pain being well known because of adult stroke [ 36 ].

Rehabilitation therapies are recommended following pediatric stroke; however, there is little evidence that describes survivors’ experiences of these services. This study was the first to report that survivors of pediatric stroke found that both muscle-specific exercise and general exercise helped in recovery. The Royal College of Physicians national guidelines for adult stroke recommend muscle-specific exercises to improve functionality and found general exercise to help aerobic fitness, gait, and prevent regression of cognitive and functional gains after stroke as well as have positive psychological effects [ 37 ]. To our knowledge, no such research has established this for pediatric stroke. A new finding from this study was that swimming facilitated recovery. This has not been researched in relation to pediatric stroke, but a feasibility study on an introductory performance-focused swimming intervention for adult cerebral palsy found that swimming helped fatigue, physical function, and mental health [ 38 ]. Speech and language therapy is commonly used during rehabilitation, despite a lack of evidence [ 25 ]. This study provides qualitative evidence that participants found this intervention useful and recommended it to other users. Similarly, this is the first qualitative study to explore patients’ experiences of physiotherapy.

Our study found that stroke-related social restriction on a child’s life has a negative emotional effect. This is an important finding, as survivors of pediatric brain injuries, including stroke, commonly have lower levels of community activity and peer social play at school [ 39 ]. Another finding from this study was that isolation and exclusion were barriers to recovery. Denham et al [ 15 ] found that following a pediatric stroke, families feel abandoned by friends or their intimate support network. Our study further characterized this dimension as “isolation” from other survivors, friends, the wider family, and society. Bullying has also been found to be a barrier in previous studies [ 40 , 41 ]. Uncertainty about the future has also been reported [ 12 ]. This study found this is a short-term barrier, as participants in the >18-year-old group did not share this concern, instead putting the stroke behind them and being more positive about the future. Financial support is accessible after recovery from stroke. However, a new finding from this study was that there was a long-term fear of losing financial support as disability allowance was reassessed. Driving is an important indicator of independence and recovery after an adult stroke [ 42 ]. This study found that the parents and the survivors were concerned about whether they could drive after a pediatric stroke, and this concern was often expressed many years after the stroke. This was only a short-term concern, as the users were reassured by adult survivors of pediatric stroke who had driving licenses and could explain the process of learning to drive with a disability. Difficulty obtaining travel insurance has been reported so far only in the literature on adult stroke [ 43 ].

This study found that both parents and children experienced anxiety-triggering events that reminded them of the happenings of their original stroke. Although a diagnosis cannot be reached from these comments, Lehman et al [ 44 ] found increased symptoms of posttraumatic stress disorder in children and their parents, suggesting that the memory of stroke has long-lasting emotional effects on the family. Our findings support previous evidence that developing comorbidities [ 3 - 7 ], behavioral problems [ 4 , 9 ], and the bereavement process caused by the stroke [ 45 ] were factors hindering recovery. A good support network [ 13 ] and third-party support and information aided recovery [ 4 , 13 , 15 ], and the users often recommended useful resources to each other.

Strengths and Limitations

The strength of this study lies in the source of data, which is a UK-wide online community with several survivors of pediatric stroke at varying stages in their lives. First, the community facilitated the discussion between participants who had a stroke recently and users who had a stroke many years ago. This allowed unprecedented insight into longer-term factors inhibiting recovery. Second, the discussions were initiated by the participants and continued in an asynchronous way, with no time, geographical, or behavioral constraints on communication. This dynamic cannot be replicated in traditional research approaches, for example, in interviews. Finally, the population that uses the forum might include people who do not partake in traditional research studies, thereby including perspectives from an underrepresented patient population [ 23 ]. The limitations of this study are that the users may not mention all the factors that affect their recovery; therefore, our analysis may not be comprehensive. First, this may be because the users do not raise everything pertinent to the research question, the users were only active over a limited period of their recovery, or the forum was moderated, so some posts may have been removed or affected by the moderation process. Second, the authenticity of posts could not be determined. Third, the data set is individuals aged >10 years, and factors affecting recovery may have changed in the time between the posts and this analysis. PPI feedback was limited to a single individual. Finally, there is a nonactive population that reads but does not compose messages. These users may be more numerous than the registered users themselves; one study [ 46 ] reported a 26:1 ratio of lurkers for every author of a message in an online forum. This population cannot be quantified or classified.

Conclusions

This study identified novel findings regarding factors affecting recovery from pediatric stroke. Raising awareness about the lived experience of survivors of pediatric stroke and the type and impact of long-term impairments is needed in health care settings, schools, workplaces, driving centers, and travel agencies so that appropriate support and information can be provided. Recovery from a stroke is an evolving process that lasts decades. Although this study has highlighted some long-term factors impacting recovery, more research needs to be performed to further establish these as well as design interventions to alleviate these barriers. This will result in effective, long-term support for survivors. In particular, fatigue, pain, and loneliness were the physical problems that were present many decades after the stroke, and there are no effective interventions reported yet.

Acknowledgments

The authors are grateful to the Stroke Association for sharing the archived file of TalkStroke on the web with them. The authors thank their patient and public involvement representative for comments on the manuscript.

ADS was partly funded by a National Institute for Health and Care Research (NIHR) Programme Grant for Applied Research (NIHR202037). The views presented in this paper are those of the authors and not necessarily those of the National Health Service, NIHR, or the Department of Health and Social Care.

Conflicts of Interest

None declared.

  • What is childhood stroke? Stroke Association. URL: https://www.stroke.org.uk/childhood-stroke/about-childhood-stroke [accessed 2023-03-12]
  • Gordon AL, Nguyen L, Panton A, Mallick AA, Ganesan V, Wraige E, et al. Self-reported needs after pediatric stroke. Eur J Paediatr Neurol. Sep 2018;22(5):791-796. [ CrossRef ] [ Medline ]
  • Greenham M, Gordon A, Anderson V, Mackay MT. Outcome in childhood stroke. Stroke. Apr 2016;47(4):1159-1164. [ CrossRef ]
  • Dunbar M, Kirton A. Perinatal stroke. Semin Pediatr Neurol. Dec 2019;32:100767. [ CrossRef ] [ Medline ]
  • Rivella C, Zanetti A, Bertamino M, Primavera L, Moretti P, Viterbori P. Emotional and social functioning after stroke in childhood: a systematic review. Disabil Rehabil. Dec 17, 2023;45(25):4175-4189. [ CrossRef ] [ Medline ]
  • Hamner T, Shih E, Ichord R, Krivitzky L. Children with perinatal stroke are at increased risk for autism spectrum disorder: prevalence and co-occurring conditions within a clinically followed sample. Clin Neuropsychol. Jul 24, 2022;36(5):981-992. [ CrossRef ] [ Medline ]
  • Sundelin H, Söderling J, Bang P, Bolk J. Risk of autism after pediatric ischemic stroke. Neurology. May 10, 2022;98(19):e1953-e1963. [ CrossRef ]
  • Wrightson JG, Zewdie E, Kuo HC, Millet GY, Kirton A. Fatigue in children with perinatal stroke: clinical and neurophysiological associations. Dev Med Child Neurol. Feb 20, 2020;62(2):234-240. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Soufi S, Chabrier S, Bertoletti L, Laporte S, Darteyre S. Lived experience of having a child with stroke: a qualitative study. Eur J Paediatr Neurol. May 2017;21(3):542-548. [ CrossRef ] [ Medline ]
  • Amlie-Lefond C. Evaluation and acute management of ischemic stroke in infants and children. Continuum (Minneap Minn). Feb 2018;24(1, Child Neurology):150-170. [ CrossRef ] [ Medline ]
  • Mrakotsky C, Williams TS, Shapiro KA, Westmacott R. Rehabilitation in pediatric stroke: cognition and behavior. Semin Pediatr Neurol. Dec 2022;44:100998. [ CrossRef ] [ Medline ]
  • McKevitt C, Topor M, Panton A, Mallick AA, Ganesan V, Wraige E, et al. Seeking normality: parents' experiences of childhood stroke. Child Care Health Dev. Jan 15, 2019;45(1):89-95. [ CrossRef ] [ Medline ]
  • Kirton A, deVeber G. Paediatric stroke: pressing issues and promising directions. Lancet Neurol. Jan 2015;14(1):92-102. [ CrossRef ]
  • Williams TS, McDonald KP, Roberts SD, Westmacott R, Dlamini N, Tam EW. Understanding early childhood resilience following neonatal brain injury from parents’ perspectives using a mixed-method design. J Int Neuropsychol Soc. May 3, 2019;25(04):390-402. [ CrossRef ]
  • Denham AM, Wynne O, Baker AL, Spratt NJ, Turner A, Magin P, et al. "This is our life now. Our new normal": a qualitative study of the unmet needs of carers of stroke survivors. PLoS One. May 8, 2019;14(5):e0216682. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Rosewilliam S, Roskell CA, Pandyan AD. A systematic review and synthesis of the quantitative and qualitative evidence behind patient-centred goal setting in stroke rehabilitation. Clin Rehabil. Jun 25, 2011;25(6):501-514. [ CrossRef ] [ Medline ]
  • Salinas J, Sprinkhuizen SM, Ackerson T, Bernhardt J, Davie C, George MG, et al. An international standard set of patient-centered outcome measures after stroke. Stroke. Jan 2016;47(1):180-186. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • De Simoni A, Shanks A, Balasooriya-Smeekens C, Mant J. Stroke survivors and their families receive information and support on an individual basis from an online forum: descriptive analysis of a population of 2348 patients and qualitative study of a sample of participants. BMJ Open. Apr 06, 2016;6(4):e010501. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Panzarasa P, Griffiths CJ, Sastry N, De Simoni A. Social medical capital: how patients and caregivers can benefit from online social interactions. J Med Internet Res. Jul 28, 2020;22(7):e16337. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Balasooriya-Smeekens C, Bateman A, Mant J, De Simoni A. Barriers and facilitators to staying in work after stroke: insight from an online forum. BMJ Open. Apr 06, 2016;6(4):e009974. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Jamison J, Sutton S, Mant J, De Simoni A. Barriers and facilitators to adherence to secondary stroke prevention medications after stroke: analysis of survivors and caregivers views from an online stroke forum. BMJ Open. Jul 16, 2017;7(7):e016814. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Jamison J, Sutton S, Mant J, De Simoni A. Online stroke forum as source of data for qualitative research: insights from a comparison with patients' interviews. BMJ Open. Mar 30, 2018;8(3):e020133. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • De Simoni A, Shanks A, Mant J, Skelton JR. Making sense of patients’ internet forums: a systematic method using discourse analysis. Br J Gen Pract. Feb 24, 2014;64(620):e178-e180. [ CrossRef ]
  • Greenham M, Anderson V, Cooper A, Hearps S, Ditchfield M, Coleman L, et al. Early predictors of psychosocial functioning 5 years after paediatric stroke. Dev Med Child Neurol. Oct 17, 2017;59(10):1034-1041. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Stroke in childhood: clinical guideline for diagnosis, management and rehabilitation. Royal College of Paediatrics and Child Health. May 2017. URL: https:/​/www.​rcpch.ac.uk/​sites/​default/​files/​2021-02/​Stroke%20guideline%2008.​04.​19%20updated%202021.​pdf [accessed 2023-03-12]
  • Braun V, Clarke V. Using thematic analysis in psychology. Qual Res Psychol. Jan 2006;3(2):77-101. [ CrossRef ]
  • Rehabiliatation. World Health Organization. Jan 30, 2023. URL: https://www.who.int/news-room/fact-sheets/detail/rehabilitation [accessed 2023-03-12]
  • Saebo UK homepage. Saebo UK. URL: https://uk.saebo.com/ [accessed 2023-03-19]
  • HemiHelp – a history. Contact. URL: https://contact.org.uk/help-for-families/information-advice-services/hemihelp/hemiplegia-support/ [accessed 2023-03-12]
  • The Maypole project homepage. The Maypole Project. URL: https://www.themaypoleproject.co.uk/ [accessed 2023-03-12]
  • Different Strokes homepage. Different Strokes. URL: https://differentstrokes.co.uk/ [accessed 2023-03-12]
  • Mobilise homepage. Mobilise. URL: https://www.mobiliseonline.co.uk/ [accessed 2023-03-12]
  • Welcome to DIAL UK. Disability Information and Advice Line. URL: https://www.dialuk.info/ [accessed 2023-03-12]
  • Headway - the brain injury association homepage. Headway - The Brain Injury Asscoiation. URL: https://www.headway.org.uk/ [accessed 2023-03-12]
  • Ezekiel L, Field L, Collett J, Dawes H, Boulton M. Experiences of fatigue in daily life of people with acquired brain injury: a qualitative study. Disabil Rehabil. Oct 04, 2021;43(20):2866-2874. [ CrossRef ] [ Medline ]
  • Harrison RA, Field TS. Post stroke pain: identification, assessment, and therapy. Cerebrovasc Dis. Mar 5, 2015;39(3-4):190-201. [ FREE Full text ] [ CrossRef ] [ Medline ]
  • Stroke guidelines 2016. Royal College of Physicians. 2016. URL: https://www.rcplondon.ac.uk/guidelines-policy/stroke-guidelines-2016 [accessed 2023-02-12]
  • Dutia IM. The effect of performance-focused swimming training on clinical outcomes in young people with cerebral palsy who have high support needs. School of Human Movement and Nutrition Sciences, The University of Queensland. 2020. URL: https://espace.library.uq.edu.au/view/UQ:8951f19 [accessed 2024-04-01]
  • van Tol E, Gorter JW, DeMatteo C, Meester-Delver A. Participation outcomes for children with acquired brain injury: a narrative review. Brain Inj. Sep 30, 2011;25(13-14):1279-1287. [ CrossRef ] [ Medline ]
  • Kirton A, deVeber G. Life after perinatal stroke. Stroke. Nov 2013;44(11):3265-3271. [ CrossRef ]
  • Sporns PB, Fullerton HJ, Lee S, Kirton A, Wildgruber M. Current treatment for childhood arterial ischaemic stroke. Lancet Child Adolesc Health. Nov 2021;5(11):825-836. [ CrossRef ]
  • Devos H, Hawley CA, Conn AM, Marshall SC, Akinwuntan AE. Driving after stroke. In: Platz T, editor. Clinical Pathways in Stroke Rehabilitation. Cham, Switzerland. Springer; 2021.
  • Hodson T, Gustafsson L, Cornwell P. The lived experience of supporting people with mild stroke. Scand J Occup Ther. Apr 02, 2020;27(3):184-193. [ CrossRef ] [ Medline ]
  • Lehman LL, Maletsky K, Beaute J, Rakesh K, Kapur K, Rivkin MJ, et al. Prevalence of symptoms of anxiety, depression, and post-traumatic stress disorder in parents and children following pediatric stroke. J Child Neurol. Jun 2020;35(7):472-479. [ CrossRef ] [ Medline ]
  • Yehene E, Brezner A, Ben-Valid S, Golan S, Bar-Nadav O, Landa J. Factors associated with parental grief reaction following pediatric acquired brain injury. Neuropsychol Rehabil. Jan 26, 2021;31(1):105-128. [ CrossRef ] [ Medline ]
  • Burri M, Baujard V, Etter JF. A qualitative analysis of an internet discussion forum for recent ex-smokers. Nicotine Tob Res. Dec 2006;8 Suppl 1(1):S13-S19. [ CrossRef ] [ Medline ]

Abbreviations

Edited by A Mavragani; submitted 29.05.23; peer-reviewed by G Engvall; comments to author 15.01.24; revised version received 04.02.24; accepted 08.03.24; published 16.04.24.

©Charlotte Howdle, William James Alexander Wright, Jonathan Mant, Anna De Simoni. Originally published in the Journal of Medical Internet Research (https://www.jmir.org), 16.04.2024.

This is an open-access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work, first published in the Journal of Medical Internet Research, is properly cited. The complete bibliographic information, a link to the original publication on https://www.jmir.org/, as well as this copyright and license information must be included.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • HHS Author Manuscripts
  • PMC10139742

Logo of nihpa

Moving Social Model Recovery Forward: Recent Research on Sober Living Houses

Douglas l. polcin.

a Behavioral Health and Recovery Studies, Public Health Institute, Oakland, CA

Amy A. Mericle

b Alcohol Research Group, Emeryville, CA

George S. Braucht

c National Alliance for Recovery Residences, St. Paul, MN

Friedner D. Wittman

d CLEW Associates, Berkeley, CA

Social model recovery is a peer centered approach to alcohol and drug problems that is gaining increased attention. This approach is well-suited to services in residential settings and typically includes living in a shared alcohol- and drug-free living environment where residents give and receive personal and recovery support. Sober Living Houses (SLHs) are recovery residences that explicitly use a social model approach. This paper describes recent research on SLHs, including new measures designed to assess their social and physical environments. We conclude that our understanding of social model is rapidly evolving to include broader, more complex factors associated with outcomes.

Introduction

It is now well recognized that many persons with alcohol or drug problems require more than acute care interventions ( Saitz, et al., 2008 ). Mutual-help programs, such as Alcoholics Anonymous (AA), have been important because persons can stay involved as long as they wish and derive the benefits of ongoing peer support. However, some individuals live in environments that undermine their recovery efforts. Residential recovery homes are a good option for many of these individuals because they provide an abstinent living environment and peer recovery support.

Because many states do not license or monitor recovery homes, ascertaining the exact number across the U.S. is difficult. However, Mericle et al. (2022) used a variety of sources to locate 10,358 residences in the U.S. Recovery homes vary in terms of their physical settings, fees, rules, requirements for involvement in mutual help groups, staffing, structure, governance, types of services offered, relationship with formal treatment programs, and lengths of stay.

Sober Living Houses

Sober living houses (SLHs) are one type of recovery home that is particularly common in California. Relative to other types of residences, SLHs are explicit in their use of a social model approach to recovery. Conceptually, the social model perspective views addiction and recovery as occurring via a reciprocal interaction between the individual and his or her social environment ( Wright, 1990 ). To maximize the beneficial effects of SLHs, service providers create a physical setting, social environment, and shared sense of responsibility among residents that supports recovery ( Wittman, et al., 2014 ). Fundamental characteristics of the social model approach include a goal of abstinence from alcohol and illicit drugs, peer support, resident input into house decisions, and resident participation in household tasks such as cooking and cleaning. In addition, residents are typically required or strongly encouraged to attend mutual help groups such as 12-step programs and develop an individualized recovery plan. Professional clinical services are not offered on-site, but residents can pursue and are encouraged to access services in the community as needed (e.g., dental, medical, mental health, job training, etc.).

SLH operations are overseen by a house manager, who is typically a person in recovery and often a person who has lived in an SLH as a resident. House managers ensure rent and bills are paid, monitor compliance with house rules, and arrange for repairs as needed. However, there is variability in how involved managers are in supporting the residents’ recovery. Recent survey data suggest some managers spend considerable time and effort supporting resident recovery, whereas others see their role as primarily administrative ( Polcin, Mahoney, & Mericle, 2020 ). One concern from a social model perspective is that managers who focus on helping residents with recovery tend to meet with them individually rather than consider ways to increase peer support and strengthen the recovery environment in the house.

Descriptions of the history and evolution of social model recovery and their origins in California SLHs are chronicled in several publications (e.g., Polcin, 2001 ; Mericle, et al., under review ; Wittman & Polcin, 2014 ). The earliest versions of SLHs began in Los Angeles in the late 1940’s in response to housing needs among persons attending AA. Known as “twelve step” houses, they implemented a very basic version of social model recovery that required alcohol and drug abstinence, attendance at AA meetings, payment of rent, and participation in upkeep of the house. In the 1970’s publications began describing the characteristics of SLHs and used the term “social model” to describe their recovery approach ( Wittman & Polcin, 2014 ). By 1990 more publications addressed social model recovery and they expanded the theoretical conceptualization and implications for practice (e.g., Shaw & Borkman, 1990 ). The overarching shift was to view addiction and recovery from an ecological systems perspective (e.g., Bronfenbrenner, 1979 ) as interactive processes between individuals and their environments. Another way to understand the shift was articulated by Borkman (2008) in her work on self-help groups: “You alone can do it, but you can’t do it alone.” This characterization acknowledges the personal responsibility for recovery as well as the importance of mutual aid (i.e., interdependence with others). Implications for SLH service providers included a stronger focus on building recovery environments that generated peer support, experiential learning, resident empowerment, and commitment to supporting others in the household.

Identifying Social Model Services

By the late 1990’s there was increased clarity about what was meant by social model in California. However, a number of questions remained. Although most SLHs and many other types of recovery homes self-identified as using a social model approach to recovery, it was often unclear to what extent they implemented a range of social model principles. For example, if a program mandated 12-step attendance and encouraged peer support, was that sufficient to be considered a social model program? If these characteristics were part of the operations of a residence but there was also a strong emphasis on clinical and medical services, should that be considered a social model program? Could a program be considered social model if there were no mechanisms in place for resident input in management decisions even if other social model characteristics were evident? To what extent was it possible to integrate some aspects of social model but not others?

A crucial step toward informing these questions was the development of the Social Model Philosophy Scale (SMPS) by Kaskutas et al (1998) , which has versions for both residential and non-residential programs. The SMPS consists of six subscales that measure distinct aspects of social model: the physical environment, staff roles, authority base, view of substance abuse problems, governance, and community orientation. Data are collected from in-person interviews with program directors or residence managers.

One purpose of the SMPS is to provide an overall cutoff score that indicates whether a program meets criteria to be described as a true social model program. Another purpose is to use subscale scores to show areas of strength and weakness in the implementation of social model. Research has shown that some aspects of social model are more prevalent than others. For example, Mericle and colleagues (2014) studied recovery residences in Philadelphia and found wide variation of subscale scores. Most recovery home service providers rated their homes high on recovery philosophy but low on peer governance. Thus, subscale scores provide a way to assess different aspects of social model so recovery residences can more strategically address social model aspects that are limited.

Classifying Types of Recovery Homes

Social model recovery principles are used to varying degrees in diverse types of recovery programs ( Borkman, Kaskutas & Owen, 2007 ), but their use might be most widespread in peer operated recovery residences. An increasing number of recovery residences are members of the National Alliance of Recovery Residences (NARR), which provides advocacy, support, training, and standards for recovery homes across the U.S. NARR’s four levels of housing range from those that are peer run (Level I) to those that are clinically focused (Level IV). NARR and its state affiliates (e.g., the Sober Living Network in California) promote using social model recovery in all four levels of recovery residences ( NARR, 2012 ). However, SLHs (Level IIs) are the most explicit in using social model recovery as a guiding influence for their operations ( Wittman & Polcin, 2014 ). In addition, social model recovery has been studied extensively in these types of residences. For these reasons, we focus our discussion herein primarily on social model issues in SLHs although many of the issues and dynamics discussed may also apply to other types or levels of recovery residences. Although Oxford Houses operationalize many aspects of social model recovery, they self-identify as separate from social model. Being part of the larger Oxford House organization is viewed as an essential component of the recovery approach. For an analysis of the relative advantages and disadvantages of leadership in SLHs and Oxford Houses see Polcin, Mahoney, and Mericle (2020b) .

Sober Living House Outcomes

Early studies of programs using a social model approach found outcomes were similar to clinically based programs but often less expensive ( Borkman, et al., 1998 ). Currently, social model recovery is largely centered in residential recovery homes and most extensively evident in SLHs. Favorable outcomes for SLH residents of have been found in several studies. For example, Polcin et al (2010a , 2010b ) examined a broad range of residents (N=300) entering 20 SLHs. Significant, sustained improvements were found at 18-month follow-up for abstinence, frequency of substance use, arrests, mental health, and employment. Improvements were noted across a broad range of residents and two characteristics of social model recovery were associated with better outcome: involvement in 12-step programs and substance use characteristics of residents’ social networks. Although residents made improvements on psychiatric severity, higher severity was associated with worse alcohol and drug outcome ( Polcin & Korcha, 2017 ).

A separate study examining outcomes for SLH residents (N=330) who had current involvement in the criminal justice system found higher severity of problems at entry into the house but similar improvements over 12 months ( Polcin, et al, 2018 ). Higher levels of recovery capital were associated with better outcomes and an add-on motivational interviewing case management (MICM) intervention was effective in providing additional benefit for higher functioning residents ( Witbrodt, et al., 2019 ).

The current paper has paper has three aims:

  • To provide an update of recent research showing the effects of SLH social environments, architectural characteristics, and neighborhoods on resident outcomes. New measures that assess the social and physical environments in houses are described.
  • A second aim considers how SLH managers and others can use recent findings to improve services. Important questions include: How should recent research findings affect the way SLH managers think about and perform their roles? What changes and modifications should SLH providers make in response to the new research? What additional research would be helpful to house managers? To what extent should providers of other types of recovery homes consider implementing social model-based changes informed by recent research on SLHs?
  • A final aim is to discuss strategies for disseminating information about social model recovery to various stakeholders. We support recovery home organizations such as NARR and its state Affiliates mandating certification and ongoing training for SLH managers and staff in other types of recovery residences.

Measuring the Recovery Environment

Recent studies of social model recovery have gone beyond previous studies that described outcomes and identified individual predictors, such as resident involvement in 12-step groups, characteristics of their social networks, and level of psychiatric severity ( Polcin et al., 2010a , 2010b ). Using the newly developed measures described below, we are moving more toward identifying house characteristics associated with outcomes, such as the strength of social model recovery in residences ( Polcin et al., 2001 ) and architectural characteristics of the physical setting that could influence recovery ( Polcin et al., 2023 ).

Recent studies have also begun to assess the influence of the neighborhoods where SLHs are located ( Mahoney et al., 2023 ; Subbaraman et al., under review ). Examples of neighborhood characteristics being studied include resident perceptions about crime, community cohesion in the neighborhood, and availability of services (e.g., public transportation). Additional factors include more objective measures, such as economic status of the neighborhood, the proximity and density of mental health and substance use services as well as destructive influences (e.g., alcohol outlets). The following sections briefly overview of house and neighborhood factors and considerations for using these findings to improve outcomes.

Recovery Home Environment Scale

The Recovery Home Environment Scale (RHES; Polcin, Mahoney & Mericle, 2021a ) is a new measure that assesses the frequency of social model activities among recovery home residents. Although the measure is useful in a variety of recovery home settings, it was developed and assessed using SLH residents. Eight items assess resident perceptions about activities in the house that are relevant to social model recovery, including social support for recovery, integration of 12-step work into daily house interactions, general and recovery oriented helping among residents, perceptions about the effectiveness of house meetings, and the degree to which residents have input into house operations. . Each item is rated on a 5-point Likert-type scale ranging from “not at all” to “a lot.” The scale’s psychometric properties were found to be strong, including measures of factor structure, reliability, construct validity, and predictive validity. Importantly, higher levels of social model in the houses were associated with significantly better outcomes, including longer retention in the house ( Mahoney, Witbrodt, Mericle & Polcin, 2021 ), higher levels of recovery capital ( Polcin, Mahoney, Witbrodt & Mericle, 2020 ), and less substance use ( Polcin, Mahoney, & Mericle, 2021a ).

Recovery Home Architecture Scale

Important aspects of recovery houses that have been largely overlooked include characteristics of the physical environment in the home. To address this shortcoming, a recent study ( Polcin et al., 2023 ) used a sample of 41 SLHs to develop a measure of architecture, the Recovery Home Architecture Scale (RHAS). The RHAS assesses the overall architecture quality in the homes and operations related to health and safety. Data are collected using observations of the home and property and are supplemented by interviews with house managers. Using the scale, the authors assessed whether physical setting characteristics of the houses were associated with outcomes. Related to that was the question of how SLHs could use mobilize architecture and maintenance procedures to improve recovery.

The RHAS consists of six subscales measuring various aspects of architecture: house maintenance, safety and security, sociability, personal and residence identity, furnishings, and outdoor areas. A copy of the instrument is available from the first author upon request. Psychometric properties included adequate levels of reliability, factor structure, and construct validity ( Polcin et al, 2023 ). At 12-month follow-up, higher scores on the sociability subscale were associated with lower psychiatric severity ( Subbaraman et al., under review ). However, other subscales were not associated with psychiatric severity and none of the subscales were associated substance use. The overall scores consistently indicated a high level of good architecture and the limited variability of the subscale scores may have made it difficult to find associations with outcomes. It might be necessary to recruit houses with more varied levels of architecture to establish significant relationships.

Using the RHES to Enhance the Social Model Recovery Environment

Most items on the RHES have clear implications for how house managers can improve social model dynamics in recovery homes. For example, if RHES items addressing involvement in 12-step or other mutual support recovery groups are low, recovery homes might improve those scores using several strategies including requiring attendance at a minimum number of meetings per week, offering on-site meetings at the house with or without community members attending, encouraging groups of residents to attending meetings together, and discussing ways to use 12-step recovery principles to address conflicts among residents and manage personal crises. To address low scores on social interaction and peer support, houses could structure regular social and recreational outings for residents. Most important is creating a supportive social climate where senior peers who have been in the residence longer engage new residents in formal and informal house activities. Senior peers also need to role model peer support, including relationship skills and development of supportive social networks. The overall goal is creating household norms of inclusion and engagement also known as belonging or community ( Porath, 2022 ; Parker, 2018 ).

Additional activities assessed on the RHES provide guidance about other ways residents can enhance social model dynamics, particularly a sense of commitment and empowerment. Examples include active engagement in giving and receiving general and recovery-oriented help, facilitating welcoming activities, participating in phase transitions and goodbye rituals that validate each individual’s contributions to the community, and providing input into discussion of house issues during house meetings. Though not directly addressed on the RHES, sharing personal experiences about the successes and challenges of working a recovery program in the residence is an additional way to help other residents and facilitate one’s own recovery.

Using the RHAS to Enhance the Physical Setting

Because the RHAS is a new measure and data linking architectural characteristics to outcomes have been limited to improved psychiatric severity (Subbaraman, et al., 2022), most of the considerations described below are based on observations of high-quality homes shown to have good alcohol and drug outcomes ( Wittman, et al., 2014 ). The contents of the subscales have clear implications for house operations. For example, houses are likely to score higher on the RHAS to the extent that house managers have systems in place to arrange for repairs (maintenance subscale), secure the house and bedrooms during night hours, and monitor the quality of furnishings (safety subscale).

Provision of some characteristics of good architecture are best implemented when selecting sites for new SLHs. For example, service providers should select houses with good socio-petal designs that facilitate social interaction. Selection of houses that include green outdoor areas can provide additional space for informal social interaction, recreation, flower and/or vegetable gardening and outdoor meals. Efficient operation of SLHs requires finding sites that contain rooms large enough for the entire house to meet. Designs that could facilitate social isolation should be avoided. Other site selection issues could include finding spatial designs where entrees are transparent so that visitors, potential contraband, and compliance with curfews can be monitored.

Facilitating Interaction of Architecture and the Social Environment

Some of the architectural considerations discussed above can be implemented in ways that might facilitate social interaction and peer support, both of which are essential features of building a social model recovery environment ( Polcin et al, 2023 ). House managers can play important roles in making architecture work not only for smooth functioning of the household, but also the quality of the social model recovery environment. For example, house managers can enhance the social and physical characteristics of the houses by mobilizing resident involvement in activities such as cooking, cleaning, simple repairs, and upkeep of outdoor areas. It is important that the manager and senior peers articulate that these activities are essential to operating a functional household, but they are also integral to building a strong recovery community. When residents follow through with tasks, fulfill responsibilities, and receive appreciation for their efforts, there is an increased sense of connection to the resident community and commitment to their peers.

It is also important for managers to consider whether they are using spaces that can accommodate the entire household to maximum benefit. House meetings involving all residents are essential to discuss updates of house operations, administrative issues, resident accomplishments, and social activities. However, house meetings also present opportunities for house managers to enhance social model dynamics by encouraging resident input into decisions affecting the household. In addition, senior residents can be engaged in articulating how issues discussed in house meetings are related to recovery and building a strong recovery environment in the house. Other uses of large spaces that can enhance the social model environment include calling impromptu house meetings to process important issues such as relapse, major rule violations, or unplanned leaving from the house. Some houses use large spaces in the house to offer open 12-step meetings to the surrounding community. Houses also use outdoor areas for social events or barbeques that are open to the surrounding community. From this perspective, facilitating social model environments goes beyond a focus within the household to include the interactive community context emphasized by Kaskutas et al (1998) . For a description of ways that house managers can facilitate social model dynamics in recovery homes and between the home and surrounding community see Polcin et al (2014) .

Social Model Recovery Across the Spectrum of Recovery Homes

Because the aforementioned studies were conducted only in SLHs, there is a need to study social model dynamics in other types of recovery houses. For example, in houses that offer on-site recovery support and clinical services (NARR Levels III and IV) the effects of social model could be independent of services, or they could interact with services in ways that facilitate or hinder recovery. In addition, the types of services offered and how they are delivered might be important as well.

Recovery homes that offer clinical services are typically governed in a more hierarchical manner where professional staff are in positions of power. This raises a concern that residents might feel less empowered, less committed to the household, and less likely to provide input into house operations and decisions. These and other characteristics of levels III and IV houses suggest it may be more challenging to implement social model recovery in these settings. However, researchers and service providers (e.g., Polcin et al, 2014 ) have described a variety of social model strategies that may be applicable to all levels of recovery homes. Drawing on their personal experiences operating houses, conceptual considerations describing social model theory, and existing studies, the authors articulated ways of understanding the challenges residents faced and potential solutions from a social model perspective.

Whether the leadership in a recovery residence is a house manager, treatment professional, or peer leader, problems and issues can be conceptualized from a household or program perspective more consistent with social model recovery than one focused primarily on individuals. When addressing problems from a social model perspective, residents, staff, and the residence leadership jointly consider questions that lead toward mobilization and enhancement of the social model environment. Examples include, how does the recovery environment in our household exacerbate or minimize the problem? Who among us has experienced this issue and what did we find helpful? What was counterproductive? What do the current residents experiencing the problem find helpful in terms of peer support? Emotional support? Practical help? Are there ways we should modify our household to be more responsive to this issue and improve our health and safety?

We suggest engaging the issues and questions posed above into ongoing management of recovery homes represents new advances for the application of social model recovery across different levels of recovery homes. As social model moves forward, it will not be enough to require attendance at mutual help groups and compliance with house rules. Residents and providers will be challenged to use a more active approach that strategically facilitates social model recovery.

Broader Context: Neighborhoods and Recovery Oriented Systems of Care (ROSC)

There is a growing recognition among recovery homes and other substance abuse service providers that recovery is best understood within a broad context that considers “Recovery Oriented Systems of Care” (ROSC) ( Kaplan, 2008 ). The idea is that persons with substance use disorders often have multiple problems and can receive help from diverse types of peer and professional resources in the community. For example, recent studies of SLHs (e.g., Mahoney et al., 2023 ; Subbaraman, et al., under review ) showed neighborhood factors associated with favorable substance use outcomes included a higher density of substance abuse and mental health services near SLHs as well as density of 12-step groups, such as Alcoholics Anonymous.

These findings align well with other studies showing individuals more involved in AA ( Polcin et al., 2010a ) and less afflicted by psychiatric symptoms ( Polcin & Korcha, 2017 ) have better outcomes. An additional analysis looked at neighborhood correlates of recovery capital among residents and found resident perceptions of neighborhood cohesion, crime, and access to transportation were associated with higher recovery capital.

It is important to note that social model strategies can be used to encourage the use of social model principles to enhance the use of local services. For example, Polcin, Korcha, and Bond (2015) described how SLH residents with psychiatric disorders can provide support to one another in terms of managing symptoms and providing information about local mental health services. In addition to sharing practical information about where services are located and how to access them, they can also share personal experiences (i.e., experiential learning) that might help residents be better prepared for what to expect.

ROSC can also include community-based resources that can help residents find work, permanent housing, social support, medical services, and legal help. In this scenario, the scope of the social model lens zooms out to include a much broader and more diverse view. For additional examples of ways that managers can mobilize good relations with the surrounding community see Polcin et al (2014) .

Considerations for Training

Although social model is the essence of recovery in SLHs, many SLH providers have only a rudimentary understanding about its history and evolution. Too often recovery residences at all levels implement a limited version of social model that simply requires a goal of abstinence, attendance at peer mutual support groups, and participation in house maintenance activities, such as cleaning and cooking. These and other social model activities need to be better understood in terms of their relevance to the social model recovery environment and the recovery process.

We suggest knowing how, where, and why social model originated and the conceptual framework of some of the early proponents can help current SLH providers implement social model more broadly and creatively. In addition, we posit this understanding is necessary to help guide social model into the future in a manner that is informed by its origins and evolution over time. It is also necessary to understanding the extent to which social model is operating in other types of recovery homes beyond SLHs and how some modifications might be needed in some settings.

Training in social model recovery needs to be offered on a regular basis. NARR facilitates Recovery Residence Provider Learning Communities on a monthly basis. Activities include didactic presentations as well as shared learning. The importance of understanding social model dynamics is evident in in NARR’s requirement that houses demonstrate the incorporation of social model principles into their operations. To succeed in fulfilling this requirement, service providers need trainings on social model characteristics described by Borkman et al (1998) : 1) an emphasis on social and interpersonal connections as the foundation of sustainable recovery, 2) the value of experiential knowledge, 3) peer-to-peer, mutual aid and other recovery supportive environments in which wellbeing is the common bond, 4) active work in an individualized recovery program, and 5) an emphasis on peer-to-peer relationships that enhance recovery/wellness objectives.

The content of trainings should include coverage of recent advances in social model theory, practice, and research. In addition to didactic presentations, we suggest recovery home organizations develop interactive learning activities (e.g., learning communities or collaboratives) where house managers visit other houses and learn through shared experiences and observations of different homes. Experiential learning is fundamental to social model recovery, yet didactic presentations are often prioritized.

Guidelines for experiential learning among house managers could be developed to help focus these interactive activities on implementation of essential elements of I social model recovery in house activities, implementation of new developments in the field, and specific issues faced by individual houses. In addition, experiential learning could expand beyond service providers to include invitations for interactions with other stakeholders, such as other service providers (mental health, medical, legal, and job training), neighbors, and local government.

Competing Demands

SLH service providers often face a host of challenges that need to be addressed if they are to survive. These include NIMBY (Not in My Back Yard) forces that resist expansion of SLH services and pressure existing houses to leave the neighborhood or reduce the number of residents. Related problems include zoning restrictions and financial pressures. In addition, many SLH managers have jobs in addition to their roles managing the houses. All of this can leave limited time for training in social model recovery or attention to building the social model environment in the house.

When manager do seek out training or informational sessions they are often on issues with direct relevance to their survival, such as legal and financial issues. In a recent paper Polcin, Mahoney, and Mericle (2020) assessed the types of training received among 35 SLH managers. The results were concerning. About two-thirds indicated they did not receive any training relevant to their house manager role over the past year. Those who did attend some type training most often reported training focused on legal and administrative issues. Training on social model recovery was reported to be rare. Not surprising, many house managers saw their roles as primarily administrative (e.g., enforce house rules, conduct intake interviews, make sure the rent and bills are paid, and arrange for needed repairs). Some managers reported spending significant amounts of time interacting with residents, (supporting their recovery, helping residents manage crises, resolving conflicts, etc.). However, these interactions appeared to be manager interactions with individuals, rather than manager led discussions with all the residents in the household, which would be more consistent with the social model approach to recovery which emphasizes peer support and experiential learning among residents.

There was strong support for some aspects of the social model approach to recovery among managers (e.g., abstinence, 12-step involvement, and peer support among residents). But there were few examples of how house managers facilitated social model principles in the houses, beyond requiring abstinence and sending residents to 12-step meetings.

The limited ways managers thought about social model recovery in their homes is an important finding particularly considering the research on the RHES showing that the strength of social model in recovery homes is associated with outcome. As social model research moves forward, we believe the focus will be on identifying variables that enhance social model and its effects on outcome. However, to improve recovery outcomes, SLH providers will need to be exposed to this research and find ways to integrate it into the operations of their homes. To the extent the homes are focused on surviving NIMBY and financial viability, new developments will be difficult to integrate.

Social model recovery in SLHs continues to emphasize original, core social model principles such as shared alcohol- and illicit drug-free living environments, a goal of abstinence, peer support, and involvement in mutual help groups. Over the last decade studies of SLHs have shown residents make significant improvements in terms of reducing or eliminating substance use, arrests, psychiatric problems, and unemployment. Studies of SLHs have also shown core social model principles, such as involvement in 12-step groups and social networks that support abstinence are associated outcome. However, as social model moves forward, we are beginning to understand social model environments from a more nuanced and complex perspective.

Recent studies have created new measures (i.e., the RHES and RHAS) designed to assess characteristics of social and physical environments of SLHs and their relationships with outcomes. While this work has only recently begun, it represents a shift in focus that may help service providers better understand the social model environment and maximize the most crucial elements. However, for these types of studies to have an impact, the effective dissemination of information to providers and other stakeholders is required. The current paper provides considerations for dissemination of new study findings and highlights the critical importance of experiential sharing of new knowledge among house managers and residents. Sharing experiences of implementing new research findings in SLHs will be vital to advancing the field.

The current paper focused on social model recovery in SLHs because these houses are the most explicit in their adoption of the social model approach to recovery. However, integration of social model principles exists to varying degrees across all four levels described by NARR (2018).

Generic strategies purported to enhance social model dynamics in houses across all four NARR levels have been described by Polcin et al (2014) . However, most current suggestions are based on provider experiences and conceptual considerations. While these are essential, studies that link characteristics of social model recovery (e.g., the RHES and RHAS) to outcomes in different types of recovery residences are needed.

As social model research and theory moves forward, it will be important to consider the mechanisms of how it promotes recovery at different time points. While individuals still reside in SLHs, the daily encounters and connections they have with other residents, the support, and the giving and receiving of help within the household may be paramount. However, research suggests most residents sustain their improvements after they leave the house ( Polcin et al., 2010a ). Understanding this transition could further strengthen long term outcomes.

It seems probable that part of what successful residents do when they leave SLHs is to reestablish aspects of social model in their post recovery home life. They attend 12-step or other types of mutual support meetings, seek out alcohol- and illicit drug-free living environments, and build prosocial networks that support recovery. They may also carry aspects of social model into their post-residence lives that are less obvious but equally impactful. Examples include internal recovery capital assets that residents acquired during their time in the SLH, such as self-confidence, self-efficacy, empowerment, spirituality, citizenship, and purpose in life. From this perspective, social model influences move beyond the boundaries of the residence and benefit previous residents and their communities. Examining these transitions and how they play out for different residents and their communities represents critically important new directions for social model research.

This article was supported by the National Institute on Drug Abuse (Grant Number DA042938) and the National Institute on Alcohol Abuse and Alcoholism (Grant Number AA028252). The funding organizations had no role in the design and conduct of the study; collection, management, analysis, and interpretation of the data; preparation, review, or approval of the manuscript; and decision to submit the manuscript for publication.

Declaration of Conflicting Interests: The author(s) declare no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

  • Borkman T.(2008). Self-help groups as participatory action. In Handbook of community movements and local organizations (pp. 211–226). Springer, Boston, MA. 10.1007/978-0-387-32933-8_14. [ CrossRef ] [ Google Scholar ]
  • Borkman T, Kaskutas LA, & Owen P.(2007). Contrasting and converging philosophies of three models of alcohol/other drugs treatment: Minnesota model, social model, and addiction therapeutic communities . Alcoholism Treatment Quarterly , 25 ( 3 ), 21–38. 10.1300/J020v25n03_03. [ CrossRef ] [ Google Scholar ]
  • Borkman TJ, Kaskutas LA, Room J, Bryan K, & Barrows D.(1998). An historical and developmental analysis of social model programs . Journal of Substance Abuse Treatment , 15 ( 1 ), 7–17. 10.1016/S0740-5472(97)00244-4. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bronfenbrenner U.(1979). The ecology of human development: Experiments by nature and design . Cambridge, MA: Harvard University. ISBN 0–674-22457–4 [ Google Scholar ]
  • Kaplan L.(2008). The Role of Recovery Support Services in Recovery-Oriented Systems of Care . DHHS Publication No. (SMA) 08–4315. Rockville, MD: Center for Substance Abuse Treatment, Substance Abuse and Psychiatric Health Services Administration. https://facesandvoicesofrecovery.org/wp-content/ . [ Google Scholar ]
  • Kaskutas LA, Greenfield TK, Borkman TJ, & Room JA (1998). Measuring treatment philosophy: A scale for substance abuse recovery programs . Journal of Substance Abuse Treatment , 15 ( 1 ), 27–36. 10.1016/S0740-5472(97)00246-8. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mahoney E, Karriker-Jaffe KJ, Mericle AA, Patterson D, Polcin DL, Subbaraman M, & Witbrodt J.(2023). Do neighborhood characteristics of sober living houses impact recovery outcomes? A multilevel analysis of observational data from Los Angeles County . Health & Place , 79 , 102951. 10.1016/j.healthplace.2022.102951. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mahoney E, Witbrodt J, Mericle AA, & Polcin DL (2021). Resident and house manager perceptions of social environments in sober living houses: Associations with length of stay . Journal of Community Psychology , 49 ( 7 ), 2959–2971. 10.1002/jcop.22620 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mericle AA, Howell J, Borkman T, Sanders BF & Polcin DL (under review) . Social model recovery and recovery housing . [ Google Scholar ]
  • Mericle AA, Miles J, Cacciola J, & Howell J.(2014). Adherence to the social model approach in Philadelphia recovery homes . International Journal of Self-Help & Self-Care , 8 ( 2 ). [ Google Scholar ]
  • Mericle AA, Patterson D, Howell J, Subbaraman MS, Faxio A, & Karriker-Jaffe KJ (2022). Identifying the availability of recovery housing in the U.S.: The NSTARR project . Drug and Alcohol Dependence , 230 , 109188. 10.1016/j.drugalcdep.2021.109188 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • National Association of Recovery Residences. (2012). A primer on recovery residences: FAQ . http://www.webcitation.org/6B7e01VSk . [ Google Scholar ]
  • National Association of Recovery Residences. (2018) NARR_Standard_V.3.0_release_11–2018.pdf . https://narronline.org/wp-content/uploads/2018/11/ .
  • Parker P.(2018). The art of gathering: How we meet and why it matters . New York: Riverhead. [ Google Scholar ]
  • Polcin DL (2001). Sober living houses: potential roles in substance abuse services and suggestions for research . Substance use & misuse , 36 ( 3 ), 301–311. 10.1081/ja-100102627 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, & Korcha R.(2017). Housing status, psychiatric symptoms, and substance abuse outcomes among sober living house residents over 18 months . Addictive Disorders & Their Treatment , 16 ( 3 ), 138. 10.1097/adt.0000000000000105 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Korcha RA, & Bond JC (2015). Interaction of motivation and psychiatric symptoms on substance abuse outcomes in sober living houses . Substance Use and Misuse , 50 ( 2 ), 195–204. 10.3109/10826084.2014.962055 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Korcha RA, Bond J, & Galloway G.(2010a). Sober living houses for alcohol and drug dependence: 18-month outcomes . Journal of Substance Abuse Treatment , 38 ( 4 ), 356–365. 10.1016/j.jsat.2010.02.003 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Korcha R, Bond J, & Galloway G.(2010b). Eighteen-month outcomes for clients receiving combined outpatient treatment and sober living houses . Journal of Substance Use , 15 ( 5 ), 352–366. 10.3109/14659890903531279 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Korcha R, Witbrodt J, Mericle AA, & Mahoney E.(2018). Motivational Interviewing Case Management (MICM) for persons on probation or parole entering sober living houses . Criminal Justice and Behavior , 45 ( 11 ), 1634–1659. 10.1177/0093854818784099 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Mahoney E, & Mericle AA (2021a). Psychometric properties of the Recovery Home Environment Scale . Substance Use and Misuse . 56 ( 8 ), 1161–1168. 10.1080/10826084.2021.1910710 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Mahoney E, Witbrodt J, & Mericle AA (2020). Recovery home environment characteristics associated with recovery capital . Journal of Drug Issues , 51 ( 2 ), 253–267. 10.1177/0022042620978393 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Mahoney E, & Mericle AA (2020b). House manager roles in sober living houses . Journal of Substance Use , 26 ( 2 ):151–155. DOI: 10.1080/14659891.2020.1789230 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin DL, Mahoney E, Witbrodt J, Mericle A, Subbaraman M.& Wittman F.(2023). Measuring architecture in recovery homes: Recovery Home Architecture Scale . Substance Use and Misuse , 58 ( 1 ): 103–110. 10.1080/10826084.2022.2148484 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Polcin D, Mericle A, Howell J, Sheridan D, & Christensen J.(2014). Maximizing social Model principles in residential recovery settings . Journal of psychoactive drugs , 46 ( 5 ), 436–443. 10.1080/02791072.2014.960112. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Porath C.(2022). Mastering community: The surprising ways coming together moves us from surviving to thriving . New York: Hachette. [ Google Scholar ]
  • Saitz R, Larson MJ, Labelle C, Richardson J, & Samet JH (2008). The case for chronic disease management for addiction . Journal of addiction medicine , 2 ( 2 ), 55–65. 10.1097/ADM.0b013e318166af74 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shaw S, & Borkman T.(editors). (1990). Social Model Alcohol Recovery: An environmental approach . Bridge-Focus Inc.: Burbank, CA. [ Google Scholar ]
  • Subbaraman M, Mahoney E, Witbrodt J, Mericle A, & Polcin DL (under review) . Multilevel effects of environment and neighborhood factors on sober living house resident 12-month Outcomes . [ Google Scholar ]
  • White W.(2009). Recovery Capital Scale . www.williamwhitepapers.com . [ Google Scholar ]
  • Wittman FD, Jee B, Polcin DL & Henderson D.(2014). The setting is the service: How the architecture of the sober living residence supports community-based recovery . International Journal of Self Help and Self Care 8 ( 2 ): 189–225. doi: 10.2190/SH.8.2.d [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wittman FD & Polcin DL (2014). The evolution of peer run sober housing as a recovery resource for California communities . International Journal of Self Help and Self Care 8 ( 2 ): 157–187. 10.2190/SH.8.2.c [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Witbrodt J, Polcin D, Korcha R, & Li L.(2019). Beneficial effects of motivational interviewing case management: A latent class analysis of recovery capital among sober living residents with criminal justice involvement . Drug and Alcohol Dependence , 200 , 124–132. 10.1016/j.drugalcdep.2019.03.017 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wright A.(1990). What is a social model? In: Shaw S; Borkman T, editors. Social Model Alcohol Recovery: An environmental approach . Burbank, CA: Bridge-Focus, Inc., p. 7–10. [ Google Scholar ]

Research on Optimization of CCUS Injection Production Parameters in High-Temperature Reservoirs Based on Intelligent Optimization Algorithms

  • Research Article-Petroleum Engineering
  • Published: 17 April 2024

Cite this article

  • Guodong Wang 1 ,
  • Zhiwei Hou   ORCID: orcid.org/0000-0001-8150-4371 1 &

The paper takes the Jidong Nanbu high temperature oil reservoir as the research object and establishes the comprehensive numerical model of CO 2 storage and displacement based on the CO 2 displacement mechanism, CO 2 heat recovery mechanism, and CO 2 geological storage mechanism. On this basis, the injection and production optimization model is established by combining the theory of multi-objective and single-objective optimization algorithm respectively. The objective function of the multi-objective injection and production optimization model is NPV and CO 2 heat recovery. The model is solved using the NSGAII and MOPSO algorithm. The results show that the objective function value obtained by the MOPSO algorithm is better than the NSGAII algorithm. The optimal values of NPV and CO 2 heat recovery were 4.781 × 10 8 CNY and 1.078 × 10 16  J respectively. The objective function of the single objective injection and production optimization model is the NPV. GA and DE optimization algorithms are used to solve the model. The results show that the DE algorithm gets the better objective function value than the GA algorithm. The optimal value of NPV is 3.69 × 10 10 CNY.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

recovery model research paper

Abbreviations

Represents the cumulative heat recovery of CO 2 (J)

Represents the total number of production wells

Represents the production rate of the i- th production well (kg/s)

Represents the specific heat capacity of the heat carrying medium ( \(J/({\text{kg}}\;{\text{K}})\) )

Represents the temperature of the produced fluid (K)

Represents the temperature of the injected fluid (K)

Represents the \(n\) dimensional decision variable, \(x = \left( {x_{1} ,x_{2} , \cdot \cdot \cdot x_{i} , \cdot \cdot \cdot ,x_{n} } \right)\)

Represents the i -th decision variable

Represents the n -dimensional decision space

Represents the \(p\) different constraints, \(h_{j} \left( x \right) \le 0,j = 1,2, \cdot \cdot \cdot ,p\)

Represents the multi-objective optimization problem composed of n objective functions to find the minimum or maximum

Represents the economic net present value in time T , CNY

Represents the CO 2 heat recovery in time T (J)

Represents the cumulative oil production of the i producing well in time T (m 3)

Represents the cumulative CO 2 production of the i producing well in time T (m 3)

Is the cumulative CO 2 storage stock in time T (t)

Represents the cumulative CO 2 injection of the i injecting well in time T (m 3 )

Represents the price of crude oil, CNY/m 3

Represents the cost of producing CO 2 gas treatment, CNY/m 3

Represents the CO 2 emission tax (CNY/ t )

Represents the cost of CO 2 injection (CNY/m 3 )

Represents the economic benefits of CO 2 thermal recovery (CNY/kw h)

Represents the discount rate

Represents the total number of producing wells

Represents the total number of injecting wells

Represents the daily heat quantity (J/day)

Represents the daily injection amount of CO 2 (m 3 /day)

Represents the minimum and maximum daily CO 2 injection rates respectively, m 3 /day

Represents the formation pressure (MPa)

Represents the minimum and maximum values of formation pressure respectively (MPa)

Represents the daily liquid production volume for oil reservoirs (m 3 /day)

Represents the minimum and maximum daily liquid production volume of the reservoir respectively (m 3 /day)

Represents the economic net present value in time T, CNY

Represents the velocity of particle i at time t

Represents the velocity of i particles at time t  + 1

Represents the position of the i -th particle at time t

Represents the position of the i -th particle at time t  + 1

Represents the acceleration constant of a particle

Represents the acceleration factor of the particle, which is the random number evenly distributed between [0, 1]

Represents the historical optimal position value of the particle and the global optimal position respectively

Represents the population obtained after mutation

Represents the population algebra

Represents the scaling factor

Represents the three individuals randomly selected from the initial population

Represents the population obtained after crossing

Represents the random number of [0, 1]

Represents the j -th component of the individual

Represents the crossover probability

Represents the random number of \([1, \cdots ,N]\)

Represents the fitness of \({\varvec{U}}_{i}^{t + 1}\) 和 \({\varvec{x}}_{i}^{t}\)

Guo, X.; Jin, Y.; Zi, J.: A 3D modeling study of effects of heterogeneity on system responses in methane hydrate reservoirs with horizontal well depressurization. Gas Sci. Eng. 115 , 205001 (2023)

Article   Google Scholar  

Godec, M.; Kuuskraa, V.; Leeuwen, T.V., et al.: CO 2 storage in depleted oil fields: the worldwide potential for carbon dioxide enhanced oil recovery. Energy Procedia 4 , 2162–2169 (2011)

Mohamed, G.R.; Jalal, F.; Davood, Z., et al.: CO 2 storage potential during CO 2 enhanced oil recovery in sandstone reservoirs. J. Petrol. Sci. Eng. 66 , 233–243 (2019)

Google Scholar  

Hu, Y.; Hao, M.; Chen, G., et al.: CO 2 displacement and burial technology and practice in Chin. Petrol. Explor. Dev. 46 (04), 716–727 (2019)

Kneafsey, T.J.; Pruess, K.: Laboratory flow experiments for visualizing carbon dioxide induced. Density-Driven Brine Convect. 82 , 123–139 (2010)

Fatima, S.; Khan, H.M.M.; Tariq, Z.; Abdalla, M.; Mohamed, M.: An experimental and simulation study of CO 2  sequestration in an underground formations; impact on geomechanical and petrophysical properties. In: SPE Middle East Oil & Gas Show and Conference (2021)

IPCC. Climate Change 2015: Synthesis Report [R]. Switzerland: IPCC (2015)

Aminu, M.D.; Nabavi, S.A.; Rochelle, C.A., et al.: A review of developments in carbon dioxide storage. Appl. Energy 208 (15), 1389–1419 (2017)

Hamid, R.J.; Zhang, D.: Optimization of carbon dioxide sequestration and enhanced oil recovery in oil reservoir. SPE133594 (2010)

Zhang, F.; Reynolds, A.C.: Optimization algorithms for automatic history matching of production data. In: ECMOR VIII - 8th European Conference on the Mathematics of Oil Recovery (2002)

Xu, Z.; Yan, B.; Gudala, M.; Tariq, Z.: A robust general physics-informed machine learning framework for energy recovery optimization in geothermal reservoirs. In: SPE EuropEC - Europe Energy Conference featured at the 84th EAGE Annual Conference & Exhibition, Vienna, Austria (2023)

Yan, B.; Xu, Z.; Gudala, M.; Tariq, Z.; Finkbeiner, T.: Reservoir modeling and optimization based on deep learning with application to enhanced geothermal systems. In: SPE Reservoir Characterisation and Simulation Conference and Exhibition, Abu Dhabi, UAE (2023)

Yan, B.; Xu, Z.; Gudala, M.; Tariq, Z.; Sun, S.; Finkbeiner, T.: Physics-informed machine learning for reservoir management of enhanced geothermal systems. Geoenergy Sci. Eng. 234 , 212663 (2024)

Gudala, M.; Tariq, Z.; Govindarajan, S., et al.: Fractured geothermal reservoir using CO 2 as geofluid: numerical analysis and machine learning modeling. ACS Omega 9 (7), 7746–7769 (2024)

Li, C.; Yan, B.; Kou, R.; Gao, S.: Rapid inference of reservoir permeability from inversion of traveltime data under a fast marching method-based deep learning framework. SPE J. 28 , 2877–2897 (2023)

Sen, M.K.; Datta-Gupta, A.; Stoffa, P.L.; Lake, L.W.; Pope, G.A.: Stochastic reservoir modeling using simulated annealing and genetic algorithm. SPE Form. Eval. 10 (01), 49–56 (1995)

Romero, C.E.; Carter, J.N.; Gringarten, A.C.; Zimmerman, R.W.: A modified genetic algorithm for reservoir characterisation. In: International Oil and Gas Conference and Exhibition in China, Beijing, China (2000)

Ouenes, A.; Bhagavan, S.: Application of simulated annealing and other global optimization methods to reservoir description: myths and realities. In: SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana

Abdassah, D.; Mucharam, L.; Soengkowo, I.; Trikoranto, H.; Sumantri, R.: Coupling seismic data with simulated annealing method improves reservoir characterization. In: SPE Asia Pacific Oil and Gas Conference, Adelaide, Australia

Wei, F.; Luo, N.: Improved intelligent optimization algorithm based on multi-level surrogate model. Control Eng. 24 (01), 83–88 (2017)

Ampomah, W.; Balch, R.; Will, R.; Cather, M.; Gunda, D.; Dai, Z.: Co-Optimization of CO 2 -EOR and storage processes under geological uncertainty. Energy Procedia 114 , 6928–6941 (2017)

Han, Z.: Advances in kriging model and surrogate optimization algorithms. J. Aeronaut. 37 (11), 3197–3225 (2016)

Kamali, F.; Hussain, F.; Cinar, Y.: A laboratory and numerical simulation study of co-optimizing CO 2 storage and CO 2 -EOR. In: SPE Asia Pacific Oil & Gas Conference and Exhibition, Adelaide, Australia (2014)

Pamukçu, Y.Z.; Gumrah, F.: A numerical simulation study of carbon-dioxide sequestration into a depleted oil reservoir. Energy Sources Part A Recover. Util. Environ. Eff. 31 (15), 1348–1367 (2009)

Li, B.: Research on CO 2 displacement and burial effects in sandstone reservoirs. Yanshan University, Qinhuangdao (2016)

Shokrollahi, A.; Arabloo, M.; Gharagheizi, F., et al.: Intelligent model for prediction of CO 2 -reservoir oil minimum miscibility pressure. Fuel 112 , 375–384 (2013)

Fathinasab, M.; Ayatollahi, S.: On the determination of CO 2 -crude oil minimum miscibility pressure using genetic programming combined with constrained multivariable search methods. Fuel 173 , 180–188 (2016)

Abdorreza, K.T.; Sassan, H.: Application of adaptive neuro fuzzy interface system optimized with evolutionary algorithms for modeling CO 2 -crude oil minimum miscibility pressure. Fuel 205 , 34–45 (2017)

Gudala, M.; Govindarajan, S.K.; Tariq, Z.: Numerical investigations and evaluation of a puga geothermal reservoir with horizontal wells using a fully coupled thermo-hydro-geomechanical model(THM)and EDAS associated with AHP. Geoenergy Sci. Eng. 228 , 212035 (2023)

Gudala, M.; Xu, Z.; Tariq, Z.; Yan, B.; Sun, S.: Numerical investigations on induced seismicity and fracture activation in fractured geothermal reservoirs. In: SPE EuropEC - Europe Energy Conference featured at the 84th EAGE Annual Conference & Exhibition, Vienna, Austria (2023)

Gudala, M.; Tariq, Z.; Yan, B.;, Sun, S.: Numerical investigations on the doublet huff and puff technology to extract heat from the geothermal reservoirs and storing of CO 2 . In: International Petroleum Technology Conference, Bangkok, Thailand (2023)

Gudala, M.; Yan, B.; Tariq, Z.; Sun, S.: Doublet huff and puff (Dhp): a new technology towards optimum sc-co 2 sequestration with stable geothermal recovery (2023). https://doi.org/10.2139/ssrn.4568399

He, K.; Bai, M.; Hu, X.; Gao, S.: CO 2 extraction dry hot rock geothermal enhanced oil recovery technology. Mod. Chem. Ind. 38 (07), 6–9 (2018)

Wang, Y.; Wang, L.; Li, H.; Bu, X.: Thermal calculation and optimization of Ganzi geothermal power generation. J. Harbin Eng. Univ. 37 (06), 873–877 (2016)

Meng, Q.; Jiang, X.: Numerical analyses of the solubility trapping of CO 2 storage in geological formations. Appl. Energy 130 , 581–591 (2014)

Silva, P.; Ranjith, P.: A study of methodologies for CO 2 storage capacity estimation of saline aquifers. Fuel 93 , 13–27 (2012)

Bachu, S.; Gunter, W.; Perkins, E.: Aquifer disposal of CO 2 : hydrodynamic and mineral trapping. Energy Convers. Manag. 35 , 269–279 (1994)

Yen, G.G.; Leong, W.F.: Dynamic multiple swarms in multi-objective particle swarm optimization. IEEE Trans. Syst. Man Cybern. Part A Syst. Hum. 39 (4), 890–911 (2009)

Liu, B.; Wang, L.; Jin, Y.: Research progress in differential evolution algorithms. Control Decis. Mak. 22 (7), 721–729 (2007)

Deb, K.: A fast elitist multi-objective genetic algorithm: NSGA-II. IEEE Trans. Evol. Comput. 6 (2), 182–197 (2000)

Coello, C.; Pulido, G.T.; Lechuga, M.S.: Handling multiple objectives with particle swarm optimization. IEEE Trans. Evol. Comput. 8 (3), 256–279 (2004)

Holland, J.H.: Adaptation in natural and artificial systems. Ann Arbor 6 (2), 126–137 (1975)

Cui, G.; Zhang, L.; Ren, S., et al.: Geochemical reaction characteristics and burial efficiency during CO 2 flooding and storage processes in oil reservoirs. J. China Univ. Pet. (Nat. Sci. Ed.) 41 (06), 123–131 (2017)

Download references

Author information

Authors and affiliations.

PetroChina Liaohe Oilfield Exploration and Development Research Institute, Panjin, Liaoning, China

Guodong Wang, Zhiwei Hou & Li Shi

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Zhiwei Hou .

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Wang, G., Hou, Z. & Shi, L. Research on Optimization of CCUS Injection Production Parameters in High-Temperature Reservoirs Based on Intelligent Optimization Algorithms. Arab J Sci Eng (2024). https://doi.org/10.1007/s13369-024-08933-7

Download citation

Received : 23 November 2023

Accepted : 29 February 2024

Published : 17 April 2024

DOI : https://doi.org/10.1007/s13369-024-08933-7

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • High temperature reservoir
  • Injection-production parameters
  • Intelligent optimization algorithm
  • Multi-objective optimization
  • Single-objective optimization
  • Find a journal
  • Publish with us
  • Track your research

AI Index Report

Welcome to the seventh edition of the AI Index report. The 2024 Index is our most comprehensive to date and arrives at an important moment when AI’s influence on society has never been more pronounced. This year, we have broadened our scope to more extensively cover essential trends such as technical advancements in AI, public perceptions of the technology, and the geopolitical dynamics surrounding its development. Featuring more original data than ever before, this edition introduces new estimates on AI training costs, detailed analyses of the responsible AI landscape, and an entirely new chapter dedicated to AI’s impact on science and medicine.

Read the 2024 AI Index Report

The AI Index report tracks, collates, distills, and visualizes data related to artificial intelligence (AI). Our mission is to provide unbiased, rigorously vetted, broadly sourced data in order for policymakers, researchers, executives, journalists, and the general public to develop a more thorough and nuanced understanding of the complex field of AI.

The AI Index is recognized globally as one of the most credible and authoritative sources for data and insights on artificial intelligence. Previous editions have been cited in major newspapers, including the The New York Times, Bloomberg, and The Guardian, have amassed hundreds of academic citations, and been referenced by high-level policymakers in the United States, the United Kingdom, and the European Union, among other places. This year’s edition surpasses all previous ones in size, scale, and scope, reflecting the growing significance that AI is coming to hold in all of our lives.

Steering Committee Co-Directors

Jack Clark

Ray Perrault

Steering committee members.

Erik Brynjolfsson

Erik Brynjolfsson

John Etchemendy

John Etchemendy

Katrina light

Katrina Ligett

Terah Lyons

Terah Lyons

James Manyika

James Manyika

Juan Carlos Niebles

Juan Carlos Niebles

Vanessa Parli

Vanessa Parli

Yoav Shoham

Yoav Shoham

Russell Wald

Russell Wald

Staff members.

Loredana Fattorini

Loredana Fattorini

Nestor Maslej

Nestor Maslej

Letter from the co-directors.

A decade ago, the best AI systems in the world were unable to classify objects in images at a human level. AI struggled with language comprehension and could not solve math problems. Today, AI systems routinely exceed human performance on standard benchmarks.

Progress accelerated in 2023. New state-of-the-art systems like GPT-4, Gemini, and Claude 3 are impressively multimodal: They can generate fluent text in dozens of languages, process audio, and even explain memes. As AI has improved, it has increasingly forced its way into our lives. Companies are racing to build AI-based products, and AI is increasingly being used by the general public. But current AI technology still has significant problems. It cannot reliably deal with facts, perform complex reasoning, or explain its conclusions.

AI faces two interrelated futures. First, technology continues to improve and is increasingly used, having major consequences for productivity and employment. It can be put to both good and bad uses. In the second future, the adoption of AI is constrained by the limitations of the technology. Regardless of which future unfolds, governments are increasingly concerned. They are stepping in to encourage the upside, such as funding university R&D and incentivizing private investment. Governments are also aiming to manage the potential downsides, such as impacts on employment, privacy concerns, misinformation, and intellectual property rights.

As AI rapidly evolves, the AI Index aims to help the AI community, policymakers, business leaders, journalists, and the general public navigate this complex landscape. It provides ongoing, objective snapshots tracking several key areas: technical progress in AI capabilities, the community and investments driving AI development and deployment, public opinion on current and potential future impacts, and policy measures taken to stimulate AI innovation while managing its risks and challenges. By comprehensively monitoring the AI ecosystem, the Index serves as an important resource for understanding this transformative technological force.

On the technical front, this year’s AI Index reports that the number of new large language models released worldwide in 2023 doubled over the previous year. Two-thirds were open-source, but the highest-performing models came from industry players with closed systems. Gemini Ultra became the first LLM to reach human-level performance on the Massive Multitask Language Understanding (MMLU) benchmark; performance on the benchmark has improved by 15 percentage points since last year. Additionally, GPT-4 achieved an impressive 0.97 mean win rate score on the comprehensive Holistic Evaluation of Language Models (HELM) benchmark, which includes MMLU among other evaluations.

Although global private investment in AI decreased for the second consecutive year, investment in generative AI skyrocketed. More Fortune 500 earnings calls mentioned AI than ever before, and new studies show that AI tangibly boosts worker productivity. On the policymaking front, global mentions of AI in legislative proceedings have never been higher. U.S. regulators passed more AI-related regulations in 2023 than ever before. Still, many expressed concerns about AI’s ability to generate deepfakes and impact elections. The public became more aware of AI, and studies suggest that they responded with nervousness.

Ray Perrault Co-director, AI Index

Our Supporting Partners

Supporting Partner Logos

Analytics & Research Partners

recovery model research paper

Stay up to date on the AI Index by subscribing to the  Stanford HAI newsletter.

Help | Advanced Search

Computer Science > Computation and Language

Title: jamba: a hybrid transformer-mamba language model.

Abstract: We present Jamba, a new base large language model based on a novel hybrid Transformer-Mamba mixture-of-experts (MoE) architecture. Specifically, Jamba interleaves blocks of Transformer and Mamba layers, enjoying the benefits of both model families. MoE is added in some of these layers to increase model capacity while keeping active parameter usage manageable. This flexible architecture allows resource- and objective-specific configurations. In the particular configuration we have implemented, we end up with a powerful model that fits in a single 80GB GPU. Built at large scale, Jamba provides high throughput and small memory footprint compared to vanilla Transformers, and at the same time state-of-the-art performance on standard language model benchmarks and long-context evaluations. Remarkably, the model presents strong results for up to 256K tokens context length. We study various architectural decisions, such as how to combine Transformer and Mamba layers, and how to mix experts, and show that some of them are crucial in large scale modeling. We also describe several interesting properties of these architectures which the training and evaluation of Jamba have revealed, and plan to release checkpoints from various ablation runs, to encourage further exploration of this novel architecture. We make the weights of our implementation of Jamba publicly available under a permissive license.

Submission history

Access paper:.

  • HTML (experimental)
  • Other Formats

license icon

References & Citations

  • Google Scholar
  • Semantic Scholar

BibTeX formatted citation

BibSonomy logo

Bibliographic and Citation Tools

Code, data and media associated with this article, recommenders and search tools.

  • Institution

arXivLabs: experimental projects with community collaborators

arXivLabs is a framework that allows collaborators to develop and share new arXiv features directly on our website.

Both individuals and organizations that work with arXivLabs have embraced and accepted our values of openness, community, excellence, and user data privacy. arXiv is committed to these values and only works with partners that adhere to them.

Have an idea for a project that will add value for arXiv's community? Learn more about arXivLabs .

IMAGES

  1. The Recovery Model

    recovery model research paper

  2. (PDF) Conceptual framework for personal recovery in mental health

    recovery model research paper

  3. The mental health recovery model and its importance for Colombian

    recovery model research paper

  4. Recovery Model of Mental Illness: A Complementary Approach to

    recovery model research paper

  5. (PDF) The Recovery Framework in Rehabilitation and Mental Health

    recovery model research paper

  6. (PDF) Remarkable recoveries: An interpretation of recovery narratives

    recovery model research paper

VIDEO

  1. The Social Model of Recovery Across the Levels of Support

  2. Business Model Research Pitch Competition

  3. Recovery Model of Mental Health- Class

  4. Segment Anything

  5. MyRecovery.me, What is a Recovery Coach?

  6. மாதிரி ஆராய்ச்சி கட்டுரைகளில் இருந்து எவ்வாறு எழுத கற்றுக்கொள்வது? Learning

COMMENTS

  1. Recovery Model of Mental Illness: A Complementary Approach to

    Recovery is often referred to as a process, an outlook, a vision, a conceptual framework or a guiding principle. There is evidence to suggest that self-management strategies based on the recovery model may have more value than models based on physical health. An analysis of the main themes in recovery based research suggest that the dominant ...

  2. Does the scientific evidence support the recovery model?

    The recovery model is a social movement that is influencing mental health service development around the world. It refers to the subjective experience of optimism about outcome from psychosis, to a belief in the value of the empowerment of people with mental illness, and to a focus on services in which decisions about treatment are taken ...

  3. The recovery model in chronic mental health: A community-based ...

    The recovery model has been enormously influential in shaping mental health services globally over the last two decades. However, empirical research on its outcomes and psychological mechanisms is sparse. This community-based case study utilised both semi-structured qualitative interviews and quanti …

  4. The recovery model in chronic mental health: A community-based

    The recovery model has become highly influential in the provision of modern mental health services, a fact which is perhaps better understood as part of social movement to protect the rights of people with mental ill-health, rather than a shift driven by the scientific evidence (Beckwith et al., 2016). Recovery principles such as person ...

  5. Models and frameworks of mental health recovery: a scoping review of

    A scoping review was conducted to identify papers describing theories, models, and frameworks of recovery to delineate the central domains of recovery. Methods Three literature search strategies were used: electronic database searching; hand-searching of key journals; and a reference list review of included papers.

  6. Recovery-Oriented Systems of Care: A Perspective on the Past, Present

    THE PAST: 2000-2010. The concept of recovery has been pushed to the forefront of behavioral health policy and practice in the United States (and elsewhere) over the last 3 decades largely through the advocacy efforts of people with behavioral health disorders rather than through advances in the effectiveness of new psychiatric medications or an accumulating body of research on clinical ...

  7. The Recovery Model and Other Rehabilitative Approaches

    3. Rehabilitation staff who saw people's lives limited artificially by systematic hopelessness, lack of opportunity and inclusiveness, and prejudice. 4. Other recovery programs and staff, especially 12-step substance abuse and trauma recovery. 1. Mental health advocates, especially the service-user/survivor movement.

  8. Integrating Evidence-Based Practices and the Recovery Model

    Integrating the recovery model and evidence-based practices. One approach to reconciling scientific and subjective approaches to treatment was recently suggested by Munetz and Frese ( 14 ). They suggested that the traditional evidence-based approach—the "medical model"—can be compatible with the recovery model.

  9. PDF Conceptual framework for personal recovery in mental health: systematic

    Personal recovery has been defined as 'a deeply personal, unique process of changing one's attitudes, values, feelings, goals, skills and/or roles . . . a way of living a satisfying, hopeful and contributing life even with the limitations caused by illness'.1 A recovery orientation is mental health policy in most Anglophone countries.

  10. The recovery model and complex health needs: What health psychology can

    This article reviews key arguments around evidence-based practice and outlines the methodological demands for effective adoption of recovery model principles. The recovery model is outlined and demonstrated as compatible with current needs in substance misuse service provision. However, the concepts of evidence-based practice and the recovery ...

  11. An Integrated Recovery-oriented Model (IRM) for mental health services

    This descriptive paper outlines a service-wide Integrated Recovery-oriented Model (IRM) for MH services, designed to enhance personally valued health, wellbeing and social inclusion outcomes by increasing access to evidenced-based psychosocial interventions (EBIs) within a service context that supports recovery as both a process and an outcome.

  12. Recovery Model of Mental Illness: A Complementary Approach to

    SUBMIT PAPER. Indian Journal of Psychological Medicine. Impact Factor: 2.8 / 5-Year Impact Factor: 2.3 . ... Editorial. First published online April 1, 2015. Recovery Model of Mental Illness: A Complementary Approach to Psychiatric Care. K. S. Jacob View all authors and affiliations. Volume 37, Issue 2. ... Sage Research Methods Supercharging ...

  13. Recovery Model of Mental Illness: A Complementary Approach to

    Recovery is often referred to as a process, an. outlook, a vision, a conceptual framework or a guiding. principle. There is evidence to suggest that self-management. strategies based on the ...

  14. Full article: Personal recovery in psychological interventions for

    The CHIME model of personal recovery has been widely adopted as a framework to ... Previous research has found that recovery-orientated intervention may improve outcomes by ... between 1980 and April 2021; and (6) available in English. Studies were excluded if they were animal studies, review papers, discussion papers, or conference ...

  15. The recovery model and patient generated outcome measures

    The recovery model and patient generated outcome measures. Robertson discusses the notions of cure and wellness. 1 In mental health, the vision of healing as "thriving within one's life as it is" is embedded in the recovery model. We always aim to reduce distressing symptoms, but labelling people as unwell and incurable can cause its own ...

  16. Recovery as the New Medical Model for Psychiatry

    Treatment grounded in recovery principles is often viewed as not being based on the "medical model." In this Open Forum the author asserts that recovery from mental illness is entirely compatible with concepts of recovery from medical illness and with new approaches to medical treatment. Three ways of conceptualizing recovery are defined: clinical recovery, illness management, and personal ...

  17. What Is Recovery? A Conceptual Model and Explication

    This paper describes a conceptual model of recovery from mental illness developed to aid the state of Wisconsin in moving toward its goal of developing a "recovery-oriented" mental health system. In the model, recovery refers to both internal conditions experienced by persons who describe themselves as being in recovery—hope, healing, empowerment, and connection—and external conditions ...

  18. The Recovery Model in Mental Health Care

    The recovery model is a holistic, person-centered approach to mental health care. The model has quickly gained momentum and is becoming the standard model of mental health care. It is based on two simple premises: It is possible to recover from a mental health condition. The most effective recovery is patient-directed.

  19. Addiction Recovery: A Systematized Review

    Theoretical paper: Model of recovery: 5: Cano et al.( 2017) USA: Cross sectional: Recovery capital: 546 participants: Questioner: 6: Duffy & Baldwin (2013) UK: Qualitative: Recovery factors: ... Although recovery research varies based on how the term has been defined and measured, there is an argument over the application of the term; recovery ...

  20. Conceptualizing eating disorder recovery research: Current perspectives

    How we research eating disorder (ED) recovery impacts what we know (perceive as fact) about it. In this paper we aim to provide an overview of the ED field's current perspectives on recovery, discuss how our methodologies shape what is known about recovery, and suggest a broadening of our methodological "toolkits" in order to form a more complete picture of recovery.

  21. Recovery Model of Mental Health Research Papers

    This paper takes a critical look at the evidence for the effectiveness of a 'Recovery' approach amongst Support, Time and Recovery workers and their clients in community mental health. The research process is discussed in the context of the attempt to give greater training and status to previously 'unqualified' community workers, and changes in ...

  22. Moving Social Model Recovery Forward: Recent Research on Sober Living

    The current paper focused on social model recovery in SLHs because these houses are the most explicit in their adoption of the social model approach to recovery. ... As social model research and theory moves forward, it will be important to consider the mechanisms of how it promotes recovery at different time points. While individuals still ...

  23. Dlr: Adversarial Examples Detection and Label Recovery for Deep ...

    Then we propose a new generative classifier and apply it to the adversarial example label recovery task. The proposed method is named as Detection and Label Recovery (DLR) defense method, which consists of Detector and Recover. Detector feeds the legitimate and adversarial examples to the target model and Recover, respectively.

  24. Journal of Medical Internet Research

    Factors influencing recovery were divided into short-term and long-term factors. Results: There were 425 posts relating to 52 survivors of pediatric stroke. Some survivors of stroke posted for themselves, while others were talked about by a third party (mostly parents; 31/35, 89% mothers). ... Journal of Medical Internet Research 8304 articles ...

  25. Moving Social Model Recovery Forward: Recent Research on Sober Living

    Sober Living Houses (SLHs) are recovery residences that explicitly use a social model approach. This paper describes recent research on SLHs, including new measures designed to assess their social and physical environments. We conclude that our understanding of social model is rapidly evolving to include broader, more complex factors associated ...

  26. Research on Optimization of CCUS Injection Production ...

    The paper takes the Jidong Nanbu high temperature oil reservoir as the research object and establishes the comprehensive numerical model of CO 2 storage and displacement based on the CO 2 displacement mechanism, CO 2 heat recovery mechanism, and CO 2 geological storage mechanism. On this basis, the injection and production optimization model is established by combining the theory of multi ...

  27. PDF Leave No Context Behind: Efficient Infinite Context Transformers with

    The model achieves even better perplexity when trained with 100K sequence length. A 1B LLM naturally scales to 1M sequence length and solves the passkey retrieval ... unified text-to-text transformer.The Journal of Machine Learning Research, 21(1):5485-5551, 2020. Nir Ratner, Yoav Levine, Yonatan Belinkov, Ori Ram, Omri Abend, Ehud Karpas, Amnon

  28. AI Index Report

    The AI Index report tracks, collates, distills, and visualizes data related to artificial intelligence (AI). Our mission is to provide unbiased, rigorously vetted, broadly sourced data in order for policymakers, researchers, executives, journalists, and the general public to develop a more thorough and nuanced understanding of the complex field ...

  29. [2403.19887] Jamba: A Hybrid Transformer-Mamba Language Model

    We present Jamba, a new base large language model based on a novel hybrid Transformer-Mamba mixture-of-experts (MoE) architecture. Specifically, Jamba interleaves blocks of Transformer and Mamba layers, enjoying the benefits of both model families. MoE is added in some of these layers to increase model capacity while keeping active parameter usage manageable. This flexible architecture allows ...