What Makes a Phenomenological Study ...

27 downloads 169 Views 194KB Size Report
Thomas, 2005); (b) discussions of phenomenology vs. ..... McAllister, Hogan, & Thomas, 2006, p. ..... more value just because it occurs more times” (Holloway.
Advancing Qualitative Methods

What Makes a Phenomenological Study Phenomenological? An Analysis of Peer-Reviewed Empirical Nursing Studies

Qualitative Health Research 20(3) 420­–431 © The Author(s) 2010 Reprints and permission: http://www. sagepub.com/journalsPermissions.nav DOI: 10.1177/1049732309357435 http://qhr.sagepub.com

Annelise Norlyk1 and Ingegerd Harder1 Abstract This article contributes to the debate about phenomenology as a research approach in nursing by providing a systematic review of what nurse researchers hold as phenomenology in published empirical studies. Based on the assumption that presentations of phenomenological approaches in peer-reviewed journals have consequences for the quality of future research, the aim was to analyze articles presenting phenomenological studies and, in light of the findings, raise a discussion about addressing scientific criteria. The analysis revealed considerable variations, ranging from brief to detailed descriptions of the stated phenomenological approach, and from inconsistencies to methodological clarity and rigor. Variations, apparent inconsistencies, and omissions made it unclear what makes a phenomenological study phenomenological. There is a need for clarifying how the principles of the phenomenological philosophy are implemented in a particular study before publishing. This should include an articulation of methodological keywords of the investigated phenomenon, and how an open attitude was adopted. Keywords phenomenology; qualitative methods; research, design Phenomenology has become a dominant philosophy that guides knowledge generation in nursing, and the term phenomenology is frequently used in nursing literature (Moi & Gjengedal, 2008; Woodgate, Ateah, & Secco, 2008). However, to adopt phenomenology as a framework for conducting nursing research is challenging. Some of the difficulties are related to understanding a complex philosophy and others are related to deciding how to accomplish a phenomenological study (Caelli, 2000, 2001; Giorgi 1997, 2006b). Phenomenology is primarily a philosophy rather than a scientific research method (Dahlberg, Dahlberg, & Nyström, 2008; Giorgi, 1997, 2006b). There are a number of schools of phenomenology, and even though they have some commonalities, they also have distinct features such as different purposes and different approaches to data analysis (Moran, 2000). There is considerable diversity, and 18 different forms of phenomenology have been identified (Caelli, 2000). This complexity challenges nurse researchers to reconsider the goals of phenomenological research and the methods they choose to achieve those goals. In the late 1990s, phenomenological research in nursing was under serious attack, with some nurse researchers being accused of disregarding fundamental principles of phenomenology and misunderstanding key concepts (Crotty, 1996; Paley, 1997). Since then, much attention has been directed toward clarification in an effort to enable

nurse researchers to pursue empirical research within the phenomenological tradition. Three main themes can be identified in this debate: (a) arguments over philosophical interpretation (Caelli, 2001; Dahlberg, 2006; Lawler, 1998; Thomas, 2005); (b) discussions of phenomenology vs. other qualitative approaches (Starks & Trinidad, 2007) and descriptive vs. interpretive approaches to doing phenomenology, focusing on commonalities and differences (Dahlberg & Dahlberg, 2004; Lopez & Willis, 2004; Rapport & Wainwright, 2006; Wojnar & Swanson, 2007); and (c) discussions of criteria to express the rigor of phenomenological nursing research for descriptive and interpretive approaches, respectively (Beck, Keddy, & Cohen, 1994; De Witt & Ploeg, 2006). Overall, the debate has focused on elaboration of theoretical complexities of phenomenology as a research approach in nursing. This article adds to the debate by providing a systematic review of what nurse researchers themselves hold as phenomenology in empirical studies. Based on the assumption that presentations of empirical phenomenological 1

University of Aarhus, Aarhus, Denmark

Corresponding author: Annelise Norlyk, University of Aarhus, Faculty of Health Sciences, Department of Nursing Science, Hoegh-Guldbergs Gade 6A, Aarhus 8000, Denmark Email: [email protected]

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

421

Norlyk and Harder approaches in peer-reviewed journals have consequences for the conduct of future research and direct implications for the legitimacy of nursing science, our review explicates details and variations within and across articles. The intention is to offer an avenue to move the debate forward and point to the possibility of constructing a minimum set of scientific criteria to be considered before publishing.

Critique of Empirical Phenomenological Nursing Research In 1996, Michael Crotty, a lecturer in education and research studies at the Faculty of Health Sciences at Flinders University (Adelaide) in Australia, published the text, Phenomenology and Nursing Research, in which he criticized nurse researchers for misinterpreting and misusing the methodology of phenomenology. Based on a review of 30 randomly selected nursing research articles dealing with phenomenology, Crotty concluded that the studies described were not phenomenological according to the European tradition, but a North American hybrid which he considered less critical, less descriptive, and subjective (Crotty, 1996). At the same time, Paley, a senior lecturer at the University of Stirling, United Kingdom, accused nurse researchers of misunderstanding key concepts of Husserlian phenomenology—i.e., phenomenological reduction, phenomenon, and essence—resulting in research that was not consistent with the original philosophy of Husserl (Paley, 1997). Likewise, he asserted that nurse researchers had misinterpreted Heidegger’s being-in-the-world and, instead, had derived a new Cartesian split between experience and reality (Paley, 1998). Nurse researchers’ responses to both Crotty’s and Paley’s critiques focused on whether the critique and its tone was dismissive of nursing research, or whether Crotty’s and Paley’s arguments offered an opportunity to critically examine phenomenological nursing research methods (Barkway, 2001; Benner, 1996; Caelli, 2000; Darbyshire, Diekelmann, & Diekelmann, 1999; Dowling, 2007; McNamara, 2005). Clearly, Crotty’s and Paley’s observations have forced nurse researchers to consider the complexities of phenomenology as a research approach. Giorgi also responded to the critique raised by Crotty and Paley, and refuted many of their arguments. He asserted that Crotty and Paley failed to make a distinction between philosophical and scientific phenomenology, by using universal philosophical criteria to determine faults with scientific procedures (Giorgi, 2000a, 2000b). However, Giorgi also admitted that there were many poor examples of the application of phenomenology in the nursing literature, and he recommended paying attention to those aspects of phenomenology that survive the mediation steps required to make the research both scientifically

and phenomenologically valid (2000b). Giorgi encouraged nurses to continue their attempt to do phenomenological research, but to do it better through a deeper understanding of the philosophy, so that a better application of the approach could ensue (2000a). In our study, we turned directly to the phenomenological literature, without any preestablished criteria, to see what the articles themselves could reveal about the application and presentation of the stated phenomenological approach. The aim was to systematically review presentations of phenomenology as a research approach and, in light of our findings, raise a discussion about possible scientific criteria that should be addressed before a study article is accepted for publication. We did not attempt to evaluate the quality of the individual study or whether what was presented in the article had actually taken place.

Study Design The critique of phenomenological nursing research was raised in 1996 and 1997, and we decided to analyze peerreviewed articles published 10 years later. This time frame was considered sufficient to identify and discuss issues of importance and move the debate forward. The sample consisted of published empirical studies covering a 1.5-year period (January, 2006, to June, 2007), with nurse researchers listed as first or corresponding author, and limited to publications in English. The search included electronic databases (CINAHL Plus, PubMed, Scopus, PsychInfo), Internet qualitative research resources, and journal-specific electronic searches. MeSH terms/Headings and search terms exploded, used singly, and/or in combination were qualitative research, nursing research, methodology, method, phenomenology, lifeworld, and lived experience. A total of 88 articles met the inclusion criteria, representing 41 periodicals. All articles were included in the initial phase of the study, and 37 were chosen for further analysis. The analysis consisted of two steps: (a) initial reading of all 88 articles, without predefined criteria, to get an impression of the authors’ presentations of the stated phenomenological approach and of the choices the authors made regarding research design; and (b) a careful and thoughtful reading of a selected group of articles using a guide developed for this second step of the analysis. The reading guide was based on the initial readings of all 88 articles, which had revealed many variations of what constitutes a phenomenological approach and what is of importance. In developing the guide we were also mindful of the critique regarding lack of rigor when researchers present phenomenological methodology and keywords. Thus, we organized the systematic review around the authors’ ways of conveying the research approach of their choice. When categorizing our material, we carefully resolved minor differences by referring to text segments

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

422

Qualitative Health Research 20(3)

Table 1. Four Groups of Phenomenological Approaches and Primary References Phenomenological (46 articles)

Phenomenological-hermeneutical (12 articles)

Hermeneutic-phenomenological (16 articles)

Benner Benner Benner van Manen van Manen van Manen Colaizzi   Colaizzi Giorgi   Giorgi Polkinghorne   Polkinghorne Ricoeur Ricoeur

across the individual article and comparing how we both understood and used the guide. We benefited from the rigor and systematic way of organizing a comprehensive review, and drew directly on the analytic component of metamethod, which refers to the study of rigor and epistemological soundness of the research methods used in previous research (Paterson, Thorne, Canam, & Jillings, 2001). We paid special attention to language use, design, findings, and validity, inspired by a reading guide developed by Sandelowski and Barroso (2002a) on how to ask questions about previous studies. The following reading guide was constructed for the systematic review: 1. Authors’ presentation of the stated phenomenological approach (e.g., philosophical ground, methodological keywords) 2. Authors’ presentation of the design and the analysis (e.g., purpose/research question, investigated phenomenon, sampling procedure and data collection, analysis process, and researcher role) 3. Authors’ presentation of the findings (e.g., explication of the phenomenon) 4. Authors’ presentation of the justification of the study (e.g., criteria addressed)

Ethical Considerations We want to make clear that the motive for this article was not to fault authors or reviewers, but to attempt to clarify the problems encountered in this type of research and to provide rationale for developing standards for submitting research identified as phenomenological. In the discussion section we carefully give several examples by referring to specific articles for illustration. This is done to make it possible for readers to follow how we came to our conclusions.

Initial Analysis The first step of the analysis, based on all 88 articles, revealed considerable variations in the description of

Interpretive-phenomenological (14 articles) Benner van Manen

the research approach and inconsistency in the use of terms, both within and across articles. Some studies were identified by the authors as phenomenological, whereas the analyses were presented as hermeneutic (Richards & Hubbert, 2007; Zust, 2006); another study was labeled hermeneutic-phenomenological, whereas the analysis was presented as phenomenological (Jonsson & Halabi, 2006). Also, the terms phenomenological approach and hermeneutic approach were used within the same abstract (Donnelly & Wiechula, 2006). There were considerable variations in the descriptions of methodological keywords, in the descriptions of the analysis process, and in overall clarity. After repeated readings, the 88 articles were grouped into four categories, according to the stated approach in the abstract and/or to the keywords listed; 46 articles were stated to be phenomenological, 12 were referred to as phenomenological-hermeneutical, 16 were called hermeneuticphenomenological, and 14 were identified by the authors as interpretive phenomenological. The initial analysis showed that although the labeling of the approaches was different across articles, the primary references could be quite similar. For example, Benner and van Manen were referred to as primary sources in all four groups, Giorgi and Colaizzi were referred to in two groups (phenomenological and hermeneuticphenomenological), and Ricouer was referred to in both the phenomenological and the phenomenological-hermeneutical groups (see Table 1). There was also considerable variation within each of the four groups regarding the primary reference for the stated research approach. The phenomenological group— i.e., the studies referred to as phenomenological by the authors—was the largest, and here we identified 13 different primary references (see Table 2). Based on the initial analysis and the identified complexity, we refrained from an in-depth analysis of all 88 articles and instead selected the largest group of 46 articles. This review is therefore limited to the articles in which the research approach was stated by the authors to be phenomenological.

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

423

Norlyk and Harder Table 2. Primary Source Referred to by the Authors in the 46 Studies Labeled as Phenomenological Primary Source

Number of Articles

Giorgi Colaizzi van Manen Pollio Benner Dahlberg Karlsson Ricoeur Drew Grounded Theory Heidegger Patton Polkinghorne No reference

10 9 7 4 2 2 2 2 1 1 1 1 1 1

Analysis of Studies Labeled Phenomenological To enable the analysis to move forward we excluded studies in which other research traditions appeared to have been used, one referring to grounded theory and one to anthropological analysis method. We also excluded articles in which both the terms phenomenological approach and hermeneutic or interpretive approach appeared, and one article in which a synthesis of intentionality was discussed as a method for disclosing a researcher’s constitutive part in phenomenological research. Thus, the findings of our review are based on 37 articles from the phenomenological group of 46, representing 20 periodicals.

Phenomenological Approach and Methodological Keywords The description of the phenomenological approach varied from a detailed description of the ontological and epistemological point of departure and how methodological considerations were implemented in the particular study (Friberg, Andersson, & Bengtsson, 2007) to brief or almost no descriptions of these aspects. The authors of 12 of the 37 articles described their research approach as descriptive phenomenology (Arthur et al., 2006; Esbensen, Swane, Hallberg, & Thome, 2008; O’Leary & Thorwick, 2006; Waite, 2006). Sixteen articles had an explicitly stated philosophical grounding—14 with reference to Husserl and/ or Merleau-Ponty, and two with reference to Heidegger. In one article building on Heidegger, the author referred to Giorgi with regard to methodical framework (Morgan, 2006), whereas others referred to Giorgi with regard to methodical framework built on Husserl (Rydeman & Törnkvist, 2006).

Most of the authors clearly considered phenomenology to be about experience. The authors of one article, however, described phenomenological research more broadly, as “in-depth study of a specific phenomenon, group or individuals, or of perceptions of social phenomena” (Notter & Burnard, 2006, p. 151). The description of methodological keywords varied from detailed descriptions related to the particular study (Friberg et al., 2007) to a one-sentence description (Heuer & Lausch, 2006; Mcilfatrick, Sullivan, & Mckenna, 2006; Spencer, 2006). The most frequently listed keywords, and those seen by the authors as central concepts in phenomenological philosophy, were experience or lived experience, bracketing, essence, and phenomenon (Esbensen et al., 2008; Kvåle, 2007; O’Leary & Thorwick, 2006). In eight articles, however, lived experience was the only mentioned keyword with explicit reference to phenomenology (Arthur et al., 2006; Joolaee, Nikbakht-Nasrabadi, Parsa-Yekta, Tschudin, & Mansouri, 2006; Newton, 2007; Prouty, Ward-Smith, & Hutto, 2006). Experience. What is meant by experience ranged from an emphasis on individual subjects, e.g., “an experience as understood by those having it” (Mcilfatrick et al., 2006, p. 297), to experiences related to a phenomenon, e.g., “the phenomenon of preparation” (Henricson, Berglund, Maatta, & Segesten, 2006, p. 241). Authors of one article drew attention to the lack of clarification of the term experience (Ruth-Sahd & Tisdell, 2007). Bracketing. The terms bracketing, reduction, and epoché seem to have been used synonymously. For example, reduction was described as “to bracket the researcher’s pre-understanding of the phenomenon such as experiences, ideas and prejudices, in order to meet the phenomenon with an ‘open mind’” (Esbensen et al., 2008, p. 3). Bracketing was described as avoiding “influence” (Karlsson, Johansson, & Lidell, 2006, p. 165) due to researcher’s preunderstandings, and epoché was described as setting “aside past associations, understandings and biases” (Kendall, 2006, p. 1152) to look and see things as if for the first time. But there were also other understandings of these terms. In a study building on Heidegger, the term phenomenological reduction was understood literally: “Originally, we had 21 categories, which were collapsed through the process of phenomenological reduction” (Ruth-Sahd & Tisdell, 2007, p. 122). There were also differences in when bracketing, as an attitude, was applied. Some authors described using bracketing throughout the research process (Hass, Coyer, & Theobald, 2006; McKeown, 2007; Tanyi, Werner, Recine, & Sperstad, 2006; Wilson, 2007), whereas others had used bracketing during the analysis process (Oh, 2006; Sunvisson, 2006; Wade, 2006), and sometimes it was used in the research process but not described (Kvåle, 2007; Logan, HackbuschPinto, & De Grasse, 2006; Waite, 2006).

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

424

Qualitative Health Research 20(3)

Essence. Although the term essence, or essential meaning, was mentioned in many of the articles, only a few of the authors elaborated on the meaning of the term and again, did so in various ways. Essence was described as both something “universal in the experience” (Kendall, 2006, p. 1152) and something situated by being related to a specific group, such as that the essence “illuminated the experience of elderly persons diagnosed with cancer” (Esbensen et al., 2008, p. 5). Essence was also equated with lived experience, as in “seeking to understand the essence or lived experience” (Notter & Burnard, 2006, p. 150). Finally, essence was related to both phenomena and experience: “the essence, i.e., the deeper meaning of the phenomenon” (Karlsson et al., 2006, p. 161); and “essence is the central underlying meaning of the experience” (Woodgate, 2006, p. 76). Phenomena. The term phenomenon typically appeared in the descriptions of methodology and analysis processes. However, most often just the term phenomenon was mentioned, and not the investigated phenomenon (Kvåle, 2007; McKeown, 2007; Morgan, 2006; Ruth-Sahd & Tisdell, 2007). In several studies the term phenomenon was not mentioned at all (Arthur et al., 2006; Bent & Magilvy, 2006; Chou, Chen, Kuo, & Tzeng, 2006; O’Leary & Thorwick, 2006), whereas one author described the term as “the building blocks of human science and the basis for all knowledge” (Newton, 2007, p. 38).

Design and Analysis Purpose/research question. Most of the authors were explicit about their attempt to explore lived experience; however, what was searched for varied. Some authors had searched for description or meaning of a phenomenon; e.g., “the phenomenon of being understood” (Shattell, McAllister, Hogan, & Thomas, 2006, p. 235). Other authors had searched for description of experience; e.g., “to access the depth of personal experiences, as well as internally interpreted meanings of those experiences” (Bent & Magilvy, 2006, p. 449). Some of the stated research purposes seemed not clearly amenable to a phenomenological approach, by focusing on explanation, comparison, or solution of problems. For example, one article included several references to phenomenology as a research method, although the purpose was stated as, “if cancer patients . . . want to talk about their feelings and difficult emotions regarding the disease and their future while they are inpatients, and if not, to identify the reason why” (Kvåle, 2007, p. 4). Another purpose was to compare caregivers’ experiences of outpatient hospital chemotherapy with their experiences of inpatient care, and the findings were presented by adopting a theoretical framework (Mcilfatrick et al., 2006). Being

explicit about employing a descriptive phenomenological approach, another author presented the purpose as “to examine what nurses can do to provide spiritual care, from the perspective of women” (Tanyi et al., 2006, p. 533). One study included three different aims, of which experience was not a part—one being to investigate whether increased knowledge and skills influenced practice (Spencer, 2006). Sampling procedure. Variation among participants was sought for in many studies with reference to phenomenological research. However, the meaning of the term variation differed. In 10 articles variation was related to demographic characteristics (Karlsson et al., 2006; Logan et al., 2006; Shattell et al., 2006). In other articles variation was related to both demographic information and different experiences in relation to the investigated phenomenon, such as stage and progress of the patients’ disease (Esbensen et al., 2008; Susleck et al., 2007; Waite, 2006). With reference to phenomenology, one author argued that the only necessary criterion is to select participants who have a lived experience of the phenomenon (Woodgate, 2006), whereas other authors stressed specific selection criteria to achieve variation. These specific criteria could be based on earlier findings (e.g., demographic information from a questionnaire; Karlsson et al., 2006), different levels of quality of life scores (Esbensen et al., 2008), or information from the participants’ medical records (Oh, 2006); or selection criteria were developed from considering the literature (McKeown, 2007; Notter & Burnard, 2006). For example, Notter and Burnard (2006) excluded men with reference to a study revealing differences in perceptions between men and women. Variation was also related to the number of participants. Fifty participants were included to achieve “as wide a cross section of participants as possible” (Notter & Burnard, 2006, p. 150), and a homogenous demographic sample was seen as a limitation of the study (Logan et al., 2006; Prouty et al., 2006). The number of participants ranged from 1 (Sunvisson, 2006) to 76 (Newton, 2007). Six, as well as 10 participants were seen as a limitation of the study (Mcilfatrick et al., 2006; Morgan, 2006). The term saturation was used as a criterion in relation to the amount of data (Kendall, 2006; Logan et al., 2006; Rydeman & Törnkvist, 2006). The term selection bias was presented as a problem in relation to voluntary participation (Rydeman & Törnkvist, 2006). Data collection. A researcher attitude of openness during data collection was stressed with reference to phenomenology. Again, there were considerable variations in the ways openness was described and the ways openness was adopted during the specific data collection procedure. In-depth interviews, as well as semistructured interviews and focus-group interviews, had been used.

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

425

Norlyk and Harder Some authors stressed the importance of an open attitude during the interview, and had used in-depth interviews (Friberg et al., 2007; Johansson & Ekebergh, 2006; Shattell et al., 2006). Other authors related openness to openended questions and had used semistructured interviews (Logan et al., 2006; Morgan, 2006). Several authors did not mention openness (Heuer & Lausch, 2006; Joolaee et al., 2006; Notter & Burnard, 2006; Mcilfatrick et al., 2006; Spencer, 2006), and had used focused questions (Mcilfatrick et al., 2006), an explicit theoretical framework as a guide for interview questions (Heuer & Lausch, 2006), or an interview guide with topics derived from the literature (Spencer, 2006). In one article the authors acknowledged that “focus-group interviews are not always thought to be compatible with a phenomenological approach” (Rydeman & Törnkvist, 2006, p. 1306). Another author had conducted a focus group interview after completion of individual interviews, and argued that it validated and generated deeper insight (Wilson, 2007). In two studies, data collection had been ongoing and concurrent with data analysis (Callister & Cox, 2006; Woodgate, 2006), whereas the other authors described analysis having taken place after data collection. Analysis process and researcher role. The descriptions of the analysis process and the researcher roles varied from a few lines (Chou et al., 2006; Heuer & Lausch, 2006; Prouty et al., 2006; Spencer, 2006) to detailed descriptions (Esbensen et al., 2008; Friberg al., 2007; Susleck et al., 2007). The terms essence, essential meaning, and general structure were used to express the focus of the analysis (Esbensen et al., 2008; Henricson et al., 2006; Kendall, 2006; Susleck et al., 2007). The term narrative structure of the phenomenon was also used (Hass et al., 2006; Logan et al., 2006). Often, the meanings of these terms were neither explicitly described nor clearly implied in the presentations of the analysis processes. When addressing the analysis, some authors focused on subjective individual experience (e.g., to describe the essence of the patients’ experiences; Morgan, 2006; Wilson, 2007; Woodgate, 2006), and others focused on the investigated phenomenon (e.g., finding the essential meaning of the phenomenon; Esbensen et al., 2006; Johansson & Ekeberg, 2006). Some authors were not explicit about whether the focus during analysis was on the phenomenon or the subjects (Arthur et al., 2006; Joolaee et al., 2006; Logan et al., 2006). In the presentation of the analysis process some authors stressed the importance of openness with respect to phenomenology; for example, using the terms bracketing, reduction, and epoché. Others did not mention any of these terms, nor did they mention openness (e.g., Logan et al., 2006; Mcilfatrick et al., 2006; Joolaee et al., 2006; Spencer, 2006), even if the study was labeled as descriptive

phenomenology or was conducted within the tradition of Husserl (Arthur et al., 2006; Bent & Magilvy, 2006; Shattell et al., 2006). The presentations of how the authors had adopted open attitudes varied from detailed description to no description. One author stated that he did not use bracketing, without further elaboration (Morgan, 2006); others described how they had striven to use bracketing (Esbensen et al., 2008; Halter, Kleiner, & Hess, 2006; Rydeman & Törnkvist, 2006), or stated that they had used bracketing, but without elaboration (Kvåle, 2007; Notter & Burnard, 2006). Some authors argued that it is necessary to set aside preunderstandings or preconceptions (Henricson et al., 2006; Kendall, 2006). Others argued, however, that it is impossible to set aside preunderstanding and that preunderstanding is a necessary condition for understanding (Johansson & Ekebergh, 2006). Instead, they described a critical attitude in which they questioned their preunderstanding to minimize its influence and to slow down the process of understanding. Likewise, continuous reflection and self-questioning were described as ways to deal with preunderstanding (Waite, 2006). Some authors used the terms reflexivity or reflexive bracketing to achieve openness, and described how this was done in the analysis (Friberg et al., 2007; Wilson, 2007). There were also variations in the ways analysis had been conducted in studies with the same primary reference. For example, with reference to Giorgi, some authors stated that the discussion was a part of the analysis (Kendall, 2006; Notter & Burnard, 2006), whereas others described a synthesis of all data as the final phase and discussed the findings afterward (Esbensen et al., 2008; Henricson et al., 2006; Karlsson et al., 2006).

Presentation of Findings We identified two distinct ways of explicating findings: focusing on the subjective experience and focusing on the phenomenon. Some authors described the essence of the experience (Halter et al., 2006; Wilson, 2007; Woodgate, 2006), and others described the essence or structure of the phenomenon (Johansson & Ekebergh, 2006; Wade, 2006; Wahlin, Ek, & Idvall, 2006). The presentation of an essence also varied, from a few lines (Henricson et al., 2006) to a long section (Karlsson et al., 2006). Authors of one article presented a main essence (divided into three subessences) substantiated by the use of many quotations (Esbensen et al., 2008), whereas others described the essence without use of quotations (Henricson et al., 2006; Karlsson et al., 2006). Presentation of findings ranged from describing a structure of themes and subthemes to consecutive descriptions using one of two headings—“Results” or “Findings.”

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

426

Qualitative Health Research 20(3)

Often, findings were presented as isolated themes rather than being explicitly interrelated. Some themes were related to a diagnosis; e.g., “the loop ileostomy” (Notter & Burnard, 2006) or “causes of diabetes” (Heuer & Lausch, 2006). Other themes were related to meaning, e.g., “seeking connection” (Logan et al., 2006). In several articles themes were substantiated by many quotations and less description, and the themes functioned as headings for these quotations, but often it was not clear whether a quotation was intended to convey a theme, a recurring feature, or to unify elements (Arthur et al., 2006; Heuer & Lausch, 2006; Kvåle, 2007). There were references to the frequency with which themes arose in a given study, or the number of participants who explicitly related having a particular experience, such as “one said,” “seven said,” or “many said” (Logan et al., 2006; McKeown, 2007; Spencer, 2006). It is not clear whether these ways of quoting participants were intended to illustrate the overall findings, or whether the authors specifically related findings to small or significant portions of the participants. Other authors explicitly presented a thematic structure from a participant perspective (Bent & Magilvy, 2006; Sunvisson, 2006; Susleck et al., 2007; Wade, 2006). There was also variation in the presentation of findings, even though the primary references were the same. For example, referring to Giorgi, one author presented the findings following headings based on the information asked of participants, and ended with “a structural description of the meaning of the phenomenon” by summing up, e.g., “Nurses are against non-disclosure” (Kendall, 2006, p. 1154). Other findings were presented as themes related to the subjective experience (Logan et al., 2006; Notter & Burnard, 2006; O’Leary & Thorwick, 2006), or as a description of an essence of the phenomenon derived from constituents (Esbensen et al., 2008; Henricson et al., 2006; Karlsson et al., 2006). One author illustrated the essential meaning by referring to meaning units and afterward using these to identify participants’ reasons for doing what they did (Kvåle, 2007). Thus, despite referring to one primary source—Giorgi, in this case—there was considerable variation in the ways findings were presented.

Justification of the Study Some authors devoted no space to justification of the study (Arthur et al., 2006; Spencer, 2006), or provided a few lines with reference to general criteria for qualitative research (Heuer & Lausch, 2006; Notter & Burnard, 2006; O’Leary & Thorwick, 2006) by mentioning that necessary revisions were made to insure confirmability and credibility (Heuer & Lausch, 2006). Others devoted more space to this issue than to the description of the analysis process (Kendall, 2006). However, what was emphasized varied

considerably. The arguments could be divided into two groups: arguments about scientific criteria to achieve trustworthiness and ensure rigor with reference to Lincoln and Guba (1985), and arguments with reference to consistency with philosophical underpinnings of phenomenology. Although reference was made to Lincoln and Guba’s (1985) criteria for credibility, dependability, confirmability, and transferability (Bent & Magilvy, 2006; Callister & Cox, 2006; Kendall, 2006; Morgan, 2006; Ruth-Sahd & Tisdell, 2007), the description of how these criteria were fulfilled varied. For example, credibility was achieved by returning the transcriptions to the participants (Morgan, 2006; Ruth-Sahd & Tisdell, 2007; Waite, 2006), by ensuring that an adequate amount of data was collected (Kendall, 2006), or by conducting the interviews in the participants’ homes (Esbensen et al., 2008). With reference to trustworthiness or rigor, some authors described that the findings were discussed with other researchers (Bent & Magilvy, 2006; Henricson et al., 2006; Shattell et al., 2006). Interdisciplinary discussions were underlined as positive (Esbensen et al., 2008; Karlsson et al., 2006); however, it was not made explicit why or how interdisciplinary discussions enhanced quality. Authors of one article stressed that agreement should be reached (Karlsson et al., 2006), whereas authors of another stressed that the purpose was not to achieve consensus but to offer competing explanations (Susleck et al., 2007). Member validation was frequently described in relation to validity, rigor, and trustworthiness. Authors of 11 articles sent identified themes or findings to participants for verification (Callister & Cox, 2006; Hass et al., 2006; Joolaee et al., 2006; Mcilfatrick et al., 2006; Tanyi et al., 2006). Some authors described rigor in relation to the characteristics of their engagement with study participants (Callister & Cox, 2006; Tanyi et al., 2006). The term saturation was also used in relation to justification of findings (Chou et al., 2006; Esbensen et al., 2008; Waite, 2006), e.g., “when no more essences of the phenomenon showed in the interviews, it was taken as a sign of saturation” (Esbensen et al., 2008, p. 5). Some authors were explicit about being true to principles of phenomenological inquiry. However, the description of these principles varied, e.g., having used purposeful sampling, secured variation among the participants, and/ or having followed a specific method rigorously (Esbensen et al., 2008; Kendall, 2006; Morgan, 2006). Others argued with reference to the researcher’s role with regard to openness, for example by referring to phenomenological terms like employing bracketing (Hass et al., 2006). To justify their study, some authors described that they engaged in the process of openness throughout the study (Friberg et al., 2007; Henricson et al., 2006; Johansson & Ekebergh, 2006; Wahlin et al., 2006). Others delimited openness to comprise open-ended interview questions

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

427

Norlyk and Harder (Bent & Magilvy, 2006; Morgan, 2006). One author argued that in keeping with the tenets of phenomenology, the literature review was undertaken postanalysis to reduce influence on the interpretation of the phenomenon (Kendall, 2006), whereas others argued that it was impossible to perform a phenomenological study without knowing something about the phenomenon (Esbensen et al., 2008; Rydemann & Törnkvist, 2006).

Discussion The analysis of presentations of phenomenology as a research approach in empirical nursing studies reveals several problems. Many of the authors did not articulate clearly which approach guided the study, nor did they identify the philosophical assumptions on which their respective studies were based. A lack of clarity made it difficult for the reader to get a sense of how the knowledge produced by the study could be evaluated, trusted, and possibly utilized. A common feature in the debate about phenomenological nursing research is that the research could be strengthened by greater attention to its philosophical underpinnings (Caelli, 2001; Crotty, 1996; Dahlberg et al., 2008; Giorgi, 2000b; Paley, 1997; Thomas, 2005). Our review of the 37 articles shows this is highly relevant but not always accomplished. The term phenomenon was not mentioned in several articles, and the investigated phenomenon was not always clearly explicated. The meaning of the term essence was implied in many articles, which is problematic, since authors who elaborated on the meaning of essence did so in various ways. Also, several articles made no reference to some of the most fundamental methodological keywords, such as reduction and bracketing. Thus, we found that the researcher role of openness was not always taken into consideration, and there were examples in which the authors’ presuppositions appeared to have crept into the study; e.g., the use of focused interview questions and the exclusion of men based on the assumption of differences in perceptions between men and women. According to Giorgi (1997), who spoke from the perspective of Husserl and MerleauPonty, no work is considered to be phenomenological if some sense of the reduction is not articulated and utilized. Having examined six randomly chosen dissertations, three from the field of psychology and three from nursing, Giorgi (2006b) commented on the variations and strategies employed. He found that the basic principles of phenomenology were often cited correctly, but they were not fully understood, nor were they always implemented correctly. Similarly, we found many descriptions about what should be done, with reference to the researcher’s open attitude, but most often no clarification of how this was actually performed.

Searching for and getting help on how to apply a phenomenological approach is not easy. Giorgi repeatedly pointed out that it is crucial to distinguish between philosophical phenomenology and scientific phenomenology, arguing that scientific phenomenology is appropriate in research, because otherwise we would practice philosophy (Giorgi, 1997, 2000a, 2000b). Philosophical phenomenology aims at describing essential universal structures of a phenomenon based on reflections of experience just from oneself (Giorgi, 1997). Scientific phenomenology aims at describing a general or typical essential structure, based on descriptions of experiences from others (Giorgi, 1997; Todres, 2005). The explication of this distinction between philosophical and scientific phenomenology differed among nurse researchers. Dahlberg et al. (2008) followed Giorgi’s path. Caelli (2000, 2001) and Dowling (2007) made a distinction between philosophical and scientific phenomenology, but they argued that both approaches have value to nursing. Wojnar and Swanson (2007) were not explicit about this distinction, and they argued that a universal description of a phenomenon is the goal in descriptive phenomenological research. These differences among researchers who intended to contribute to the clarification of how to accomplish a phenomenological study make it difficult for nurse researchers to get help to carry out a phenomenological study. Hence, we found that essence was referred to both as a universal structure and as a structure typical for a specific group. We also identified problems in the 37 articles regarding presentations of the design and analysis processes. In particular, we found that purposeful sampling was often mentioned as theoretical rationale and recruitment method for the sample. Our analysis shows that it was not made clear what purposeful sampling meant in a phenomenological study. Variation within the sample was stressed by many authors; however, variation within the sample was expressed as demographic characteristics, size, and other specific criteria. The literature on phenomenological research does not offer much on sampling procedures—only sample expressions of individual lifeworld experiences relevant to the phenomenon of interest (Giorgi, 1997; Giorgi & Giorgi 2003; Todres, 2005). Instead, it is argued that the aim in a phenomenological study is to understand a phenomenon more deeply through adequate exposure to the qualities of the phenomenon that are described by those experiencing the phenomenon (Dahlberg et al., 2008; Giorgi & Giorgi, 2003; Todres, 2005). Furthermore, it is argued that variation within the phenomenological framework is variation of the experience (Todres, 2005). So, although several authors included in our study paid much attention to sampling procedures and how variation was achieved, the phenomenological literature does not. We argue that sampling criteria presented in many of the articles—such as size,

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

428

Qualitative Health Research 20(3)

cross section, and demographical information—are examples of empirical research criteria and that the authors confused these with phenomenological research criteria. One of Crotty’s (1996) and Paley’s (1997) claims was that the target of nursing phenomenological research was purely the subjective experiences of the participants, instead of unfolding phenomena. We found that when subjective experience was a theme for the analysis, there was a tendency in many of the articles to present the findings as a summary of direct descriptions of the participants’ experiences, using many quotations, instead of structural descriptions of situations to which participants were attaching meaning. Dahlberg et al. (2008, p. 255) stated that the focus in phenomenological analysis is to describe “how the phenomenon is, i.e., not what the informants said about it” (italics in original). Accordingly, the use of many quotations and less description is not considered to be within a phenomenological framework. This can also be seen as an example of misuse of quotations, which makes identifying the findings a challenging task for readers (Sandelowski & Barroso, 2002b). We also found that the findings in many articles were not explicitly interrelated, and the meanings were not related to each other. Hence, there was no presentation of a structure. Some articles stood out as an analysis of data focusing on numbers, or as placing more value on meanings that occurred often. However, within the phenomenological approach, it is argued that “a ‘part meaning’ is not given more value just because it occurs more times” (Holloway & Todres, 2005, p. 96). Scrutinizing some of the method literature referred to in the 37 articles, we found that the authors of the method literature clearly paid attention to unfolding phenomena. Furthermore, there seemed to be agreement that phenomenological analysis entails an exploration of what appears to the subject and the manner of its appearing, with the goal of elucidating the nature of the phenomenon as an essential experience (Giorgi, 1997; Todres, 2005). The frequent use of participants as evaluators was a significant finding in relation to the authors’ justifications of their respective studies. According to Giorgi, however, use of participants as evaluators of the findings within phenomenology is wholly indefensible theoretically (Giorgi, 2006a). Using participants as evaluators overlooks the fact that participants describe experience from the perspective of the natural attitude, whereas analysis is conducted from the perspective of the phenomenological attitude. These perspectives do not match, and if the final word is going to be given to the individual undergoing the experience, there are no arguments for an analysis to take place. Likewise, a new problem arises if a discrepancy exists between the researcher and participants, or between participants, in determining which perspective takes priority (Giorgi, 2006a). Giorgi argued for a greater trust

in subjectivity and the safeguards that the method and phenomenological attitude provide. It is remarkable that the authors of many of the articles reviewed in our study drew on Lincoln and Guba’s criteria (1985) to evaluate or justify their research. These criteria are generic to qualitative research, and were not specifically developed to evaluate phenomenological research. They were described 20 years ago in another historical context, where the agenda was to communicate differences between quantitative and qualitative research (Sandelowski, 2006). Recently, Sandelowski (2006) indicated that her own work from 1986—in which she used criteria from Lincoln and Guba’s work—“should be retired” (p. 645). As early as 1994, Beck and colleagues argued for refocusing the discussion of the global criteria of Lincoln and Guba toward criteria specific to the various approaches to phenomenological research, including differences between descriptive and interpretive approaches (Beck et al., 1994). This is highly relevant, especially when considered in light of the number of schools of phenomenology. Even though they have some commonalities, they also have distinct features, e.g., different purposes and different approaches to data analysis (Moran, 2000). Philosophical differences have methodological implications for empirical research. Phenomenology grounded on Husserl focuses on description of meaning of phenomena experienced using some type of bracketing to identify essences (Giorgi, 1997, 2006b). Phenomenology grounded on hermeneutic phenomenologists like Heidegger, Gadamer, and Ricoeur focuses on interpretation of meaning in which preconceptions are integrated into the research findings (Caelli, 2000; Dahlberg et al., 2008). Therefore, nurse researches are challenged to be careful when naming the research approach, and to be specific about what it entails. Our review reveals the importance of setting up a minimum of criteria that should be addressed to some degree when publishing. Giorgi stated that within descriptive phenomenology, at least three criteria must be implemented: (a) description—obtained from others from the perspective of the natural attitude, (b) adoption of the attitude of phenomenological reduction—bracketing past knowledge of the phenomenon, and (c) imaginative variation—a search for invariant meaning (essence) for a context (1997). We argue that identification of the philosophical assumptions on which the study is based must include, at a minimum, an articulation of methodological keywords and, especially, how an open attitude is adopted and maintained throughout the research process. These criteria are more relevant than details about specific sampling procedures that might not be of relevance in a phenomenological study. Clarifying phenomenological research criteria might also help the author to avoid staying at the level of subjective experiences, and instead describe the unfolded

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

429

Norlyk and Harder phenomenon. In addition, the interrelation of the parts of the investigated phenomenon might become more apparent. However, the challenge is not only to articulate the principles of phenomenological philosophy, but to ensure that they are wed to good scientific practices. As stated by Giorgi (2000b), the responsibility for adding quality to phenomenological research rests not only on researchers but on all those engaged in mediating between philosophy and scientific practice. We argue that a major responsibility also rests on supervisors, reviewers, and editors of periodicals.

Declaration of Conflicting Interests The authors declared no conflicts of interest with respect to the authorship and/or publication of this article.

Funding The authors disclosed receipt of the following financial support for the research and/or authorship of this article: Funding for this study was provided by the Novo Nordic Foundation, Aarhus University Hospital, and the Danish Nursing Council.

References

Conclusion This review of 37 articles, presenting studies identified as phenomenological by the authors, reveals considerable variation, ranging from brief to detailed descriptions of the chosen phenomenological approach, and from inconsistencies to methodological clarity and rigor. The variations, apparent inconsistencies, and omissions make it unclear what makes a phenomenological study phenomenological. Our analysis points to the need for understanding differences among phenomenological research approaches, and the need to distinguish more clearly. Turning to the phenomenological articles themselves without predetermined criteria enabled us to identify the seemingly most challenging aspects for the authors. It is not enough to refer to phenomenology as a research approach. It is important to clarify how the principles of phenomenological philosophy are implemented in the particular study. Published empirical studies based on phenomenology should include a minimum of scientific criteria such as articulation of methodological keywords, articulation of the investigated phenomenon, and description of how an open attitude was adopted throughout the research process. We considered limiting our review to studies identified by the authors as “descriptive phenomenological,” but found only 12 articles. By choosing the largest group of articles, in which the term phenomenological approach was used by the authors, we might have included some studies that were anchored in an interpretive phenomenological tradition. Thus, our identification of inconsistencies, lack of clarity, and so forth, might appear to be severe. However, the analysis revealed that there were also considerable variations within and across presentations of the studies that were specifically identified as “descriptive phenomenological.” Nurse researchers are faced with the difficulties inherent in offering adequate philosophical and methodological explication and still keeping within the word limits of journals. But there is room for improvement, and we must continue to critically examine applications of stated phenomenological approaches. This ongoing debate about the quality of phenomenological nursing research is valuable in our search for new knowledge.

Arthur, D., Drury, J., Sy-Sinda, M. T., Nakao, R., Lopez, A., Gloria, G., et al. (2006). A primary health care curriculum in action: The lived experience of primary health care nurses in a school of nursing in the Philippines: A phenomenological study. International Journal of Nursing Studies, 43(1), 107-112. Barkway, P. (2001). Michael Crotty and nursing phenomenology: Criticism or critique? Nursing Inquiry, 8(3), 191-195. Beck, C. T., Keddy, B. A., & Cohen, M. Z. (1994). Reliability and validity issues in phenomenological research. Western Journal of Nursing Research, 16(3), 254-267. Benner, P. (1996). Phenomenology and nursing research. Nursing Inquiry 3(4), 257-258. Bent, K. N., & Magilvy, J. K. (2006). When a partner dies: Lesbian widows. Issues in Mental Health Nursing, 27(5), 447-459. Caelli, K. (2000). The changing face of phenomenological research: Traditional and American phenomenology in nursing. Qualitative Health Research, 10(3), 366-377. Caelli, K. (2001). Engaging with phenomenology: Is it more of a challenge than it needs to be? Qualitative Health Research, 11(2), 273-281. Callister, L. C., & Cox, A. H. (2006). Opening our hearts and minds: The meaning of international clinical nursing electives in the personal and professional lives of nurses. Nursing & Health Sciences, 8(2), 95-102. Chou, F. H., Chen, C. H., Kuo, S. H., & Tzeng, Y. L. (2006). Experience of Taiwanese women living with nausea and vomiting during pregnancy. Journal of Midwifery & Women’s Health, 51(5), 370-375. Crotty, M. (1996). Phenomenology and nursing research. Melbourne, Victoria, Australia: Churchill Livingstone. Dahlberg, K. (2006). The essence of essences—The search for meaning structures in phenomenological analysis of lifeworld phenomena. International Journal of Qualitative Studies on Health and Well-Being, 1(1), 11-19. Dahlberg, K., & Dahlberg, H. (2004). Description vs. interpretation—A new understanding of an old dilemma in human science research. Nursing Philosophy, 5(3), 268-273. Dahlberg, K., Dahlberg, H., & Nyström, M. (2008). Reflective lifeworld research. Lund, Sweden: Studentlitteratur. Darbyshire, P., Diekelmann, J., & Diekelmann, N. (1999). Reading Heidegger and interpretive phenomenology: A response to the work of Michael Crotty. Nursing Inquiry, 6, 17-25.

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

430

Qualitative Health Research 20(3)

De Witt, L., & Ploeg J. (2006). Critical appraisal of rigor in interpretive phenomenological nursing research. Journal of Advanced Nursing, 55(2), 215-229. Donnelly, F., & Wiechula, R. (2006). The lived experience of a tracheostomy tube change: A phenomenological study. Journal of Clinical Nursing, 15(9), 1115-1122. Dowling, M. (2007). From Husserl to van Manen. A review of different phenomenological approaches. International Journal of Nursing Studies, 44(1), 131-142. Esbensen, B. A., Swane, C. E., Hallberg, I. R., & Thome, B. (2008). Being given a cancer diagnosis in old age: A phenomenological study. International Journal of Nursing Studies, 45(3), 393-405. Friberg, F., Andersson, P. E., & Bengtsson, J. (2007). Pedagogical encounters between nurses and patients in a medical ward—A field study. International Journal of Nursing Studies, 44(4), 534-544. Giorgi, A. (1997). The theory, practice, and evaluation of the phenomenological method as a qualitative research procedure. Journal of Phenomenological Psychology, 28(2), 235-261. Giorgi, A. (2000a). The status of Husserlian phenomenology in caring research. Scandinavian Journal of Caring Sciences, 14, 3-10. Giorgi, A. (2000b). Concerning the application of phenomenology to caring research. Scandinavian Journal of Caring Sciences, 14, 11-15. Giorgi, A. (2006a). Concerning variations in the application of the phenomenological method. Humanistic Psychologist, 34(4), 305-319. Giorgi, A. (2006b). Difficulties encountered in the application of the phenomenological method in the social sciences. Análise Psicológica, 24(3), 353-361. Giorgi, A., & Giorgi. B. (2003). Phenomenology. In J. L. Smith (Ed.), Qualitative psychology—A practical guide to research methods (pp. 25-51). London: Sage. Halter, M. J., Kleiner, C., & Hess, R. F. (2006). The experience of nursing students in an online doctoral program in nursing: A phenomenological study. International Journal of Nursing Studies, 43(1), 99-105. Hass, H., Coyer, F. M., & Theobald, K. A. (2006). The experience of agency nurses working in a London teaching hospital. Intensive & Critical Care Nursing, 22(3), 144-153. Henricson, M., Berglund, A. L., Maatta, S., & Segesten, K. (2006). A transition from nurse to touch therapist—A study of preparation before giving tactile touch in an intensive care unit. Intensive & Critical Care Nursing, 22(4), 239-245. Heuer, L., & Lausch, C. (2006). Living with diabetes: Perceptions of Hispanic migrant farmworkers. Journal of Community Health Nursing, 23(1), 49-64. Holloway, I., & Todres, L. (2005). The status of method: Flexibility, consistency and coherence. In I. Holloway (Ed.), Qualitative research in health care. London: Open University Press.

Johansson, A., & Ekebergh, M. (2006). The meaning of wellbeing and participation in the process of health and care— Women’s experiences following a myocardial infarction. International Journal of Qualitative Studies on Health and Well-Being, 1(2), 100-112. Jonsson, A., & Halabi, J. (2006). Work related post-traumatic stress as described by Jordanian emergency nurses. Accident and Emergency Nursing, 14(2), 89-96. Joolaee, S., Nikbakht-Nasrabadi, A., Parsa-Yekta, Z., Tschudin, V., & Mansouri, I. (2006). An Iranian perspective on patients’ rights. Nursing Ethics, 13(5), 488-502. Karlsson, A., Johansson, M., & Lidell, E. (2006). Endurance— Integration of strength and vulnerability in relatives’ response to open heart surgery as a lived experience. International Journal of Qualitative Studies on Health and WellBeing, 1(3), 159-166. Kendall, S. (2006). Being asked not to tell: Nurses’ experiences of caring for cancer patients not told their diagnosis. Journal of Clinical Nursing, 15(9), 1149-1157. Kvåle, K. (2007). Do cancer patients always want to talk about difficult emotions? A qualitative study of cancer inpatients communication needs. European Journal of Oncology Nursing, 11(4), 320-327. Lawler, J. (1998). Phenomenologies as research methodologies for nursing: From philosophy to research practice. Nursing Inquiry (5), 104-111. Lincoln, Y. S, & Guba, E. (1985). Naturalistic inquiry. Beverly Hills, CA: Sage. Logan, J., Hackbusch-Pinto, R., & De Grasse, C. (2006). Women undergoing breast diagnostics: The lived experience of spirituality. Oncology Nursing Forum, 33(1), 121-126. Lopez, K., & Willis, D. (2004). Descriptive versus interpretive phenomenology: Their contributions to nursing knowledge. Qualitative Health Research, 14(5), 726-735. Mcilfatrick, S., Sullivan, K., & Mckenna, H. (2006). What about the carers?: Exploring the experience of caregivers in a chemotherapy day hospital setting. European Journal of Oncology Nursing, 10, 294-303. McKeown, F. (2007). The experiences of older people on discharge from hospital following assessment by the public health nurse. Journal of Clinical Nursing, 16(3), 469-476. McNamara, M. (2005). Knowing and doing phenomenology: The implications of the critique of “nursing phenomenology” for a phenomenological inquiry: A discussion paper. International Journal of Nursing Studies, 42, 695-704. Moran, D. (2000). Introduction to phenomenology. London: Routledge. Morgan, R. (2006). Using clinical skills laboratories to promote theory-practice integration during first practice placement: an Irish perspective. Journal of Clinical Nursing, 15(2), 155-161. Moi, A. L., & Gjengedal, E. (2008). Life after burn injury: Striving for regained freedom. Qualitative Health Research, 18, 1621-1631.

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015

431

Norlyk and Harder Newton, S. E. (2007). Alcohol relapse and its relationship to the lived experience of adult liver transplant recipients. Gastroenterology Nursing, 30, 37-42. Notter, J., & Burnard, P. (2006). Preparing for loop ileostomy surgery: Women’s accounts from a qualitative study. International Journal of Nursing Studies, 43(2), 147-159. Oh, J. (2006). Stroke patients’ experiences of sharing rooms with dementia patients in a nursing home. International Journal of Nursing Studies, 43(7), 839-849. O’Leary, J., & Thorwick, C. (2006). Fathers’ perspectives during pregnancy, postperinatal loss. Journal of Obstetric, Gynecologic, and Neonatal Nursing, 35(1), 78-86. Paley, J. (1997). Husserl, phenomenology and nursing. Journal of Advanced Nursing, 26, 187-193. Paley, J. (1998). Misinterpretive phenomenology: Heidegger, ontology and nursing research. Journal of Advanced Nursing 27(4) 817-824. Paterson, B. L., Thorne, S. E., Canam, C., & Jillings, C. (2001). Meta-study of qualitative health research. A practical guide to meta-analysis and meta-synthesis. Thousand Oaks, CA: Sage. Prouty, D., Ward-Smith, P., & Hutto, C. J. (2006). The lived experience of adult survivors of childhood cancer. Journal of Pediatric Oncology Nursing, 23(3), 143-151. Rapport, F., & Wainwright, P. (2006). Phenomenology as a paradigm of movement. Nursing Inquiry, 13(3), 228-236. Richards, J., & Hubbert, A. O. (2007). Experiences of expert nurses in caring for patients with postoperative pain. Pain Management Nursing, 8(1), 17-24. Ruth-Sahd, L. A., & Tisdell, E. J. (2007). The meaning and use of intuition in novice nurses: A phenomenological study. Adult Education Quarterly, 57(2), 115-140. Rydeman, I., & Törnkvist, L. (2006). The patient’s vulnerability, dependence and exposed situation in the discharge process: Experiences of district nurses, geriatric nurses and social workers. Journal of Clinical Nursing, 15(10), 1299-1307. Sandelowski, M. (2006). In response to: de Witt L. & Ploeg J. (2006) Critical appraisal of rigor in interpretive phenomenological nursing research. Journal of Advanced Nursing 55(2), 215-229. Journal of Advanced Nursing, 55(5), 643-645. Sandelowski, M., & Barroso, J. (2002a). Reading qualitative studies. International Journal of Qualitative Methods, 1(1). Retrieved January 12, 2008, from http://www.ualberta.ca/~ijqm/ Sandelowski, M., & Barroso, J. (2002b). Finding the findings in qualitative studies. Journal of Nursing Scholarship, 34(3), 213-219. Shattell, M. M., McAllister, S., Hogan, B., & Thomas, S. P. (2006). “She took the time to make sure she understood”: Mental health patients’ experiences of being understood. Archives of Psychiatric Nursing, 20(5), 234-241. Spencer, R. L. (2006). Nurses’, midwives’ and health visitors’ perceptions of the impact of higher education on professional practice. Nurse Education Today, 26(1), 45-53.

Starks, H., & Trinidad, S. B. (2007). Choose your method: A comparison of phenomenology, discourse analysis, and grounded theory. Qualitative Health Research, 17, 1372-1380. Sunvisson, H. (2006). Stopped within a track: Embodied experiences of late-stage Parkinson’s disease. International Journal of Qualitative Studies on Health and Well-Being, 1(2), 91-102. Susleck, D., Willocks, A., Secrest, J., Norwood, B. K., Holweger, J., Davis, M., et al. (2007). The perianesthesia experience from the patient’s perspective. Journal of Perianesthesia Nursing, 22(1), 10-20. Tanyi, R. A., Werner, J. S., Recine, A. C., & Sperstad, R. A. (2006). Perceptions of incorporating spirituality into their care: A phenomenological study of female patients on hemodialysis. Nephrology Nursing Journal, 33(5), 532-538. Thomas, S. P. (2005). Through the lens of Merleau-Ponty: Advancing the phenomenological approach to nursing research. Nursing Philosophy, 6(1), 63-76. Todres, L. (2005). Clarifying the life-world: Descriptive phenomenology. In I. Holloway (Ed.), Qualitative research in health care. London: Open University Press. Wade, J. (2006). “Crying alone with my child”: Parenting a school age child diagnosed with bipolar disorder. Issues in Mental Health Nursing, 27(8), 885-903. Wahlin, I., Ek, A. C., & Idvall, E. (2006). Patient empowerment in intensive care—An interview study. Intensive & Critical Care Nursing, 22(6), 370-377. Waite, R. (2006). The psychiatric educational experiences of advance beginner RNs. Nurse Education Today, 26(2), 131-138. Wilson, D. W. (2007). From their own voices: The lived experience of African American registered nurses. Journal of Transcultural Nursing, 18(2), 142-149. Wojnar, D. M., & Swanson, K. M. (2007). Phenomenology. An exploration. Journal of Holistic Nursing, 25(3), 172-180. Woodgate, R. L. (2006). Living in a world without closure: Reality for parents who have experienced the death of a child. Journal of Palliative Care, 22(2), 75-82. Woodgate, R. L., Ateah, C., & Secco, L. (2008). Living in a world of our own: The experience of parents who have a child with autism. Qualitative Health Research, 18, 1075-1083. Zust, B. L. (2006). Meaning of INSIGHT participation among women who have experienced intimate partner violence. Issues in Mental Health Nursing, 27(7), 775-793.

Bios Annelise Norlyk, RN, MScN, is a doctoral student at the University of Aarhus, Institute of Public Health, Department of Nursing Science, in Aarhus, Denmark. Ingegerd Harder, RN, PHD, is an associate professor at the University of Aarhus, Institute of Public Health, Department of Nursing Science, in Aarhus, Denmark.

Downloaded from qhr.sagepub.com at University of Victoria on January 30, 2015