Logo for Pressbooks

Want to create or adapt books like this? Learn more about how Pressbooks supports open publishing practices.

Research Guides

Ethnographic Case Studies

Jeannette Armstrong; Laura Boyle; Lindsay Herron; Brandon Locke; and Leslie Smith

Description

This research guide discusses ethnographic case study. While there is much debate over what, precisely, delimits a case study , the general consensus seems to be that ethnographic case studies differ from other types of case studies primarily in their focus, methodology, and duration. In essence, ethnographic case studies are case studies “employing ethnographic methods and focused on building arguments about cultural, group, or community formation or examining other sociocultural phenomena” (Schwandt & Gates, 2018, p. 344), typically with a long duration, per the demands of ethnographic work. In essence, ethnographic case studies are case studies “employing ethnographic methods and focused on building arguments about cultural, group, or community formation or examining other sociocultural phenomena” (Schwandt & Gates, 2018, p. 344), typically with a long duration, per the demands of ethnographic work. Indeed, in its very situatedness, ethnography has a “case study character” and is “intimately related” to case studies (Ó Rian, 2009, p. 291); though there is currently a move to extract ethnographic work from overly situated contexts and use extended case methods, “[e]thnographic research has long been synonymous with case studies, typically conceived of as grounded in the local and situated in specific, well-defined and self-contained social contexts” (Ó Rian, 2009, p. 290). Because ethnography, in practice, is often a kind of case study, it’s useful to consider ethnography and case studies each in their own right for a fuller picture of what ethnographic case study entails.

Ethnographic research is one approach under the larger umbrella of qualitative research. Methodologically, it is, “a theoretical, ethical, political, and at times moral orientation to research, which guides the decisions one makes, including choices about research methods” (Harrison, 2014, p. 225), that is at its crux “based upon sharing the time and space of those who one is studying” (Ó Rian, 2009, p. 291)–a situated, nuanced exploration seeking a thick description and drawing on methods such as observation and field notes. According to …an ethnography focuses on an entire culture-sharing group and attempts to develop a complex, complete description of the culture of the group. Creswell and Poth (2018), an ethnography focuses on an entire culture-sharing group and attempts to develop a complex, complete description of the culture of the group. In doing so, ethnographers look for patterns of behavior such as rituals or social behaviors, as well as how their ideas and beliefs are expressed through language, material activities, and actions (Creswell & Poth, 2018). Yin (2016)  suggests that ethnographies seek “to promote embedded research that fuses close-up observation, rigorous theory, and social critique. [Ethnographies foster] work that pays equal attention to the minutiae of experience, the cultural texture of social relations, and to the remote structural forces and power vectors that bear on them” (p. 69).

Case study research, meanwhile, is characterized as an approach “that facilitates exploration of a phenomenon within its context using a variety of data sources” (Baxter & Jack, 2008, p. 544). The aim of case studies is precise description of reconstruction of cases (Flick, 2015). The philosophical background is a qualitative, constructivist paradigm based on the claim that reality is socially constructed and can best be understood by exploring the tacit, i.e., experience-based, knowledge of individuals. There is some debate about how to define a The philosophical background is a qualitative, constructivist paradigm based on the claim that reality is socially constructed and can best be understood by exploring the tacit, i.e., experience-based, knowledge of individuals. “case” (e.g., Ó Rian, 2009), however. As Schwandt and Gates (2018) write, “[A] case is an instance, incident, or unit of something and can be anything–a person, an organization, an event, a decision, an action, a location”; it can be at the micro, meso, or macro level; it can be an empirical unit or a theoretical construct, specific or general; and in fact, “what the research or case object is a case of may not be known until most of the empirical research is completed” (p. 341). The two authors conclude that given the multifarious interpretations of what case study is, “[b]eyond positing that case study methodology has something to do with ‘in-depth’ investigation of a phenomenon . . . , it is a fool’s errand to pursue what is (or should be) truly called ‘case study’” (p. 343, 344).

Baxter, P., & Jack, S. (2008). Qualitative case study methodology: Study design and implementation for novice researchers. The Qualitative Report, 13 (4), 544-559.

Creswell, J. W., & Poth, C. N. (2018). Qualitative inquiry & research design: Choosing among five approaches (4th ed.). Los Angeles, CA: SAGE.

Flick, U. (2015). Introducing research methodology . Los Angeles, CA: SAGE.

Rian, S. (2009). Extending the ethnographic case study. In D. Byrne & C. C. Ragin (Eds.), The SAGE handbook of case-based methods (pp. 289–306). Thousand Oaks, CA: SAGE.

Schwandt, T. A., & Gates, E. F. (2018). Case study methodology. In N. K. Dezin & Y. S. Lincoln (Eds.), The SAGE handbook of qualitative research (5th ed.; pp. 341-358). Thousand Oaks, CA: SAGE.

Yin, R. K. (2016). Qualitative research from start to finish (2nd ed.). New York, NY: The Guilford Press.

Key Research Books and Articles on Ethnographic Case Study Methodology

Fusch, G. E., & Ness, L. R. (2017). How to conduct a mini-ethnographic case study: A guide for novice researchers. The Qualitative Report , 22 (3), 923-941.  Retrieved from https://nsuworks.nova.edu/tqr/vol22/iss3/16

In this how-to article, the authors present an argument for the use of a blended research design, namely the Ethnographic Case Study, for student researchers. To establish their point of view, the authors reiterate recognized research protocols, such as choosing a design that suits the research question to ensure data saturation. Additionally, they remind their reader that one must also consider the feasibility of the project in terms of time, energy, and financial constraints.

Before outlining the benefits and components of the Ethnographic Case Study approach, the authors provide detailed narratives of ethnographic, mini-ethnographic (sometimes referred to as a focused ethnography ), and case study research designs to orient the reader. Next, we are introduced to the term mini-ethnographic case-study design, which is defined as a blended design that is bound in time and space and uses qualitative ethnographic and case study collection methods. The benefits of such an approach permit simultaneous generation of theory and the study of that theory in practice, as it allows for the exploration of causality.

Ethnographic Case Study research shares many characteristics with its parent approaches.  For example, subjectivity and bias are present and must be addressed. Next, data triangulation is necessary to ensure the collected qualitative data and subsequent findings are valid and reliable. Data collection methods include direct observation, fieldwork, reflective journaling, informal or unstructured interviews, and focus groups. Finally, the authors discuss three limitations to the ethnographic case study. First, this design requires the researcher to be embedded, yet the duration of time may not be for as long when compared to full-scale ethnographic studies.  Second, since there are fewer participants, there should be a larger focus on rich data as opposed to thick data, or said differently, quality is valued over quantity. Third, the researcher must be aware that the end-goal is not transferability, but rather the objective is to gain a greater understanding of the culture of a particular group that is bound by space and time.

Gregory, E. & Ruby, M. (2010) The ‘insider/outsider’ dilemma of ethnography: Working with young children and their families in cross-cultural contexts. Journal of Early Childhood Research, 9 (2), 1-13. https://doi.org/10.1177/1476718X10387899

This article focuses on the dilemma of insider and outsider roles in ethnographic work. It challenges the notion that a researcher can be both an insider and an outsider at the same time. There is no insider/outsider status; it is one or the other–not both.

It is easy to make assumptions about one’s status as an insider. It is not uncommon for a researcher to assume that because one is working amongst his/her “own” people sharing a similar background, culture, or faith that she/he is an insider. Likewise, a researcher may assume that it will be easy to build rapport with a community with which he/she has commonalities; however, it is important to keep in mind that the person may be an insider but the researcher may not have this same status. When the person enters into the protective space of family or community as a researcher, it is similar to being an outsider. Being a researcher makes one different, regardless of the commonalities that are shared. It is not the researcher’s presumed status of “insider” or “outsider” that makes the difference; rather, researcher status is determined by the participants or community that is being studied. It is wise for researchers to understand that they are distinctively one of “them” as opposed to one of “us”. This is not to say that researchers cannot become an “insider” to some degree. But to assume insider status, regardless of the rationale, is wrong. Assuming common beliefs across cultures or insider status can lead to difficulties that could impact the scope or nature of the study.

In conclusion, regardless of the ethnographic design (e.g., realist ethnography, ethnographic case study, critical ethnography), it is important for the researcher to approach the study as an “outsider”. Although the outsider status may change over time, it essential to understand that when one enters a community as a researcher or becomes a researcher within a community, insider status must be earned and awarded according to the participants in the community.

Ó Rian, S. (2009). Extending the ethnographic case study. In D. Byrne & C. C. Ragin (Eds.), The SAGE handbook of case-based methods (pp. 289–306). Thousand Oaks, CA: SAGE.

In this chapter, Ó Rian valorizes the problems and potential hiding within the vagaries of ethnographic “case” boundaries, arguing that “whereas the fluid and multi-faceted aspects of the ethnographic case pose dilemmas for ethnographers, they can also become resources for ethnographers in exploring theoretical and empirical questions” (p. 292). Indeed, he views the idea of firm case boundaries as a weakness, as “definitions of the case will rule in and out certain social processes,” and suggests ethnography’s flexibility can deal with this problem well because it permits researchers to “question the boundaries of the case as the study proceeds,” leading to a “de- and re-construction of the case that . . . places ethnography at the centre of a resurgent contextualist paradigm of social inquiry . . . that is increasingly self-consciously exploring its own theoretical and methodological foundations” (p. 304). Most of the chapter delves into these possibilities for exploration, offering an insightful (if occasionally difficult to follow) perspective on how they have been proceeding.

The chapter offers considerations that might be particularly helpful to researchers undertaking ethnographic case studies who are struggling to connect their cases, so firmly rooted in a particular context and their own personal experiences and observations, to a bigger picture. Ó Rian elucidates the reflexive strategies various ethnographers have adopted as they’ve sought “[t]o achieve a link between context-specific data and meso- or macro-level generalizations,” categorizing these strategies into three “interlocking extensions of case study research” (p. 292): personal extensions (related to “the shaping of the boundaries of the case by the ethnographer’s location within the field and . . . how ethnographers can convey their personalized experiences and tacit learning to readers” [p. 292]), theoretical extensions (which bridge the gap between the situated worlds being explored and “the larger structures and processes that produced and shaped them” [p. 292]), and empirical extensions (“creative efforts to experiment with the empirical boundaries of the ethnographic case” [p. 292] by bringing in, for example, historical context, social networks, etc.). The crux of his argument is that ethnographic researchers have a prime opportunity to push against the boundaries of their context and “extend their cases across space, time and institutional structures and practices” so that the ethnographer is “multiply, if perhaps a bit uncomfortably, situated” (p. 304), and also to include an “emphasis on the ongoing process of theoretical sampling within the process of the ethnographic study, with close attention to be paid to the paths chosen and rejected, and the reasons for these decisions” (p. 304). These kinds of extensions offer an opportunity for theories to “be refined or reconstructed” as the researcher attempts to locate their personal experience within a broader framework, allowing “[t]he case study . . . to challenge and reconstruct the preferred theory” while also connecting the case to a larger body of work, particularly because theory “carries the accumulated knowledge of previous studies” (p. 296).

Ó Rian’s in-depth descriptions of how other researchers have varyingly handled these personal, theoretical, and empirical extensions might be a bit overwhelming to novice researchers but overall can offer a way to “locate their cases within broader social processes and not solely within their own personal trajectories” (p. 294)–while also helping to situate their reflections and extensions within a larger body of literature replete with researchers struggling with similar questions and concerns.

This chapter offers an  in-depth, generally accessible (but occasionally overwhelming) overview of case studies of all sorts and integrates an extensive review of relevant literature. The authors provide an informed perspective on various considerations and debates in the case study field (e.g., varying definitions of what a “case” is construed to be; interpretive vs. critical realist orientations; the relative benefits of and techniques involved in different types of approaches), helping novice researchers locate and better describe their own approach within the context of the field. The information is quite detailed and delves into a wide variety of case study types, suggesting this chapter might best be first skimmed as an initial introduction, followed by more careful readings of relevant sections and perusal of the key texts cited in the chapter. The breadth of this chapter makes it a helpful resource for anyone interested in case-study methodology.

The authors do not specifically explore ethnographic case studies as a separate type of case study. They do, however, briefly touch on this idea, locating ethnography within the interpretive orientation (comprising constructivist approaches offering “phenomenological attention to lived experience” [p. 344]). The authors also cite researchers who distinguish it due to its “[employing] ethnographic methods and focus on building arguments about cultural, group, or community formation or examining other sociocultural phenomena” (p. 344). Ethnographic case study is placed in contrast to case studies that use non-ethnographic methods (e.g., studies “relying perhaps on survey data and document analysis”) or that “are focused on ‘writing culture’” (p. 344).

Two aspects of this chapter are particularly useful for novice researchers. First, it is worth highlighting the authors’ discussion of varying definitions of what a “case” is, as it can provide an interesting reconceptualization of the purpose of the research and the reason for conducting it. The second noteworthy aspect is the authors’ detailed descriptions of the four main case study uses/designs ( descriptive, hypothesis generation or theory development, hypothesis and theory testing , and contributing to normative theory ), which the authors beautifully align with the respective purposes and methods of each type while also offering insight into relevant conversations in the field.

Further Readings

Moss, P. A., & Haertel, E. H. (2016). Engaging methodological pluralism. In D. H. Gitomer & C. A. Bell (Eds.), Handbook of Research on Teaching (pp. 127–247). Washington, DC: American Educational Research Association.

Simons, H. (2014). Case study research: In-depth understanding in context. In P. Leavy (Ed.), The Oxford handbook of qualitative research (pp. 455–470). Oxford, UK: Oxford University Press.

Recent Dissertations Using Ethnographic Case Study Methodology

Cozzolino, M. (2014). Global education, accountability, and 21st century skills: A case of curriculum innovation . Retrieved from ProQuest Dissertations & Theses Global. (Order Number 3648007)

This dissertation is self-described as an ethnographic case study of a small, public, suburban high school in Pennsylvania. In this study, the researcher investigates the school’s process of integrating global education into its curriculum by implementing a school-wide initiative (Global Studies Initiative or GSI) as well as a program of study (Global Studies Credential or GSC). Cozzolino asserts that her framework has been shaped by both social constructivism and critical/Freirean pedagogy. From the constructivist view, she views knowledge as constructed through social interaction, and thus she sought to understand the world in which the research participants work, learn, and experience large parts of their lives. It is here that she situates the first three research questions that entail looking at the the GSI and the GSC in terms of their features, rationales, and implementations. The fourth question involves understanding the students’ views and perceptions of the GSC and here the author takes up a critical and Freirean pedagogy to honor and hear the voices of the students themselves.

The study design is therefore an embedded single-case study in that it is bound by the place (Olympus High School) and by its population. Furthermore, it is also a case within a case, as it seeks to understand the students’ perspectives of the global programming. The case study is ethnographically rooted through the multiple ethnographic data sources such as participant-observations and a prolonged engagement at the research site. Cozzolino embedded herself in the research site over a five-year period and became an active and invested member of the school community, thereby establishing a sound rationale for an ethnographic case-study approach.

The author concludes that there were some competing priorities about the overall initiative from stakeholders inside and outside the school district. This resulted in a less than ideal implementation of the program of study across the curriculum. Nonetheless, the students who were enrolled in these courses reported it to be a worthwhile experience. While Cozzolino presents specific recommendations for the improvements at Olympus High, she also offers implications for several other groups. First, she provides advice for implementation to other educational institutions that aim to integrate a global focus into their curriculum. Next, she gives recommendations for local, state, and national policy changes. Finally, she gives suggestions for engaging all parties in fruitful discourse to achieve their ultimate goal of implementing a meaningful and valuable global education curriculum.

Hamman, L. (2018). Reframing the language separation debate: Language, identity, and  ideology in two-way immersion . Retrieved from ProQuest Dissertations & Theses Global. (Order Number 2089463322)

This study explored the issues of surrounding language separation in two-way immersion (TWI) classrooms. The author looked at how classroom language practices and teacher ideologies influenced the student experience and how the students’ understanding of what it means to be bilingual is influenced in a classroom that purports to be equitable in terms of language use.

The study is theoretically grounded in sociocultural, critical, and postcultural theories and adapted Lemke’s ecosocial system to conceptualize TWI classroom. Hamman also drew upon translanguaging theory and dynamic bilingualism to provide a framework for a more modern and nuanced perspective of bilingualism, bilingual learning, and bilingual students.

The author combined a single-case study approach with ethnographic methods to “engage in close analysis of classroom language use and the discursive negotiation of identities and ideologies, while situating these analyses within a rich understanding of the sociolinguistic context of this TWI classroom” (p. 78-79). She employed various ethnographic methods such as taking fieldnotes, conducting participant observations, interviewing, and memoing. The study is “bound” in that it takes place in one 2nd-grade classroom with one teacher and 18 students over the course of one year.

Hamman concludes that student perspectives on language separation should be considered, since this forced separation of language influenced how they thought of their developing bilingualism and identity as bilinguals. Furthermore, the study envisages a linguistic “middle ground” to strict separation that allows for appropriate and meaningful spaces for linguistic negotiation. Finally, this dissertation asserts that the strict separation of languages codifies a monoglossic ideology mindset and limits learners’ possibilities for learning and making connections across languages.

Kim, S. (2015). Korean migrant youth identity work in the transnational social field: A link between identity, transnationalism, and new media literacy . Retrieved from University of Missouri-St. Louis Institutional Repository Library. https://irl.umsl.edu/dissertation/158/

This doctoral dissertation takes an ethnographic case study approach to explore the identity formation of transnational Korean youth. The researcher, herself a Korean immigrant to the U.S. navigating complex identity processes, focuses on these research questions: “1) what are the contexts in which migrant youth negotiate their identities? 2) how do youth understand and negotiate their sense of belonging? 3) how do youth’s [sic] cultural and literacy practices inform and shape their identities? 3i) how do youth make use of transnational new media for their identity work? 3ii) how do literacy practices potentially shape their identities?” (p. 7).

Drawing on Leander and McKim (2013), the author conceptualizes her study as a “connective ethnography” (p. 36) encompassing multiple spaces, both digital and physical, in which “space” comprises a variety of relationships, instead of a more traditional ethnography bounded by physical space. The “case study” aspect, meanwhile, refers to the four specific participants in which she chose to focus. She chose Korean immigrants in St. Louis, in general, due to their mobility between the U.S. and Korea, their high use of digital communication and information technology, and their limited access to the cultural resources of Korea in a Midwestern city. From an initial 32 possible participants purposively selected, the researcher chose four focal participants based on their Korean ethnicity, biliteracy in Korean and English, age (between 11 and 19 years old), residence in the U.S. (for at least 2 years), and their use of digital communication technologies. Data sources included an initial screening survey, an identity map each participant created, informal recorded conversations, recorded interviews in either English or Korean, field notes from the researcher’s interactions with the youth in various settings (home, school, community centers), and “literacy documents” (evidence of literacy practices from participants’ school and home, emails to the researcher, or activities in digital spaces). She used social semiotic multimodal discourse analysis and what she describes as “grounded theory thematic analysis” to analyze the data.

This is a reflective, thoughtful, and interesting dissertation. The author carefully notes the relationship between the data sources and her research questions, specifically addresses steps she took to ensure the validity of the data (e.g., triangulation via multiple data sources and theoretical frameworks, member checks, and feedback from her professors and other researchers), and discloses her own positionalities and biases. Her discussion includes not only a clear thematic exploration of her findings but also offers specific practical suggestions for how her findings can be applied and extended in the classroom.

Internet Resources

Abalos-Gerard Gonzalez , L. (2011). Ethnographic research . Retrieved from https://www.slideshare.net/lanceabalos/ethnographic-research-2?from_action=save

Created by Lance Gerard G. Abalos, teacher at the Department of Education-Philippines, this SlideShare, Ethnographic Research , explains that, regardless of specific design, ethnographic research should be undertaken “without any priori hypothesis to avoid predetermining what is observed or that information is elicited from informants . . .hypotheses evolve out of the fieldwork itself” (slide 4). It is also suggested that researchers refer to individuals from whom information is gathered as ‘informants’ is preferred over the term ‘participants’ (slide 4).

According to Abalos, “It is not the data collection techniques that determine whether the study is ethnographic, but rather the ‘socio-cultural interpretation’ that sets it apart from other forms of qualitative inquiry” (slide 6). A social situation always has three components: a place, actors, and activities (slide 8) and it is the socio-cultural interpretation of the interactions of these three that is the focus of the ethnographic research.

Ethnographic questions should guide what the researcher sees, hears, and collects as data (slide 9). When writing the ethnography, it is essential to ‘bring the culture or group to life’ through the words and descriptions used to describe the place, actors, and activities.

Abalos describes three types of ethnographic designs:

  • Realist Ethnographies : an objective account of the situation, written dispassionately from third-person point of view, reporting objectively on information learned from informants, containing closely edited quotations (slide 11-12).
  • Ethnographic Case Studies : researchers focus on a program, event, or activity involving individuals rather than a group, looking for shared patterns that develop as a group as a result of the program, event, or activity (slide 13).
  • Critical Ethnographies: incorporating a ‘critical’ approach that includes an advocacy perspective, researchers are interested in advocating against inequality and domination (slide 14).

As ethnographic data is analyzed, in any design (e.g., realist, case study, critical), there is a shift away from reporting the facts to making an interpretation of people and activities, determining how things work, and identifying the essential features in themes of the cultural setting (slide 22). “The ethnographer must present the description, themes, and interpretation within the context or setting of the culture-sharing group (slide 23).

Brehm, W. (2016, July 21). FreshEd #13 – Jane Kenway . Retrieved from http://www.freshedpodcast.com/tag/ethnography/ (EDXSymposium: New Frontiers in Comparative Education).

Jane Kenway is with the Australian Research Council and is an emeritus professor at Monash University in Melbourne, Australia. In this podcast, she explains “traditional’ forms of ethnography and multi-sited global ethnography, which are her area of specialization. She considers “traditional” ethnography to have three components: space, time, and mobility.

Insider/outsider stance is explained within the context of spatiality, community, and culture of space specific to ‘traditional” ethnography. Researchers are outsiders who are attempting to enter a space and become insiders, then leave the space once the research is completed. Research is conducted over an extended period of time in one place/space. As a result, researchers will get to know in an extremely intimate manner the ways of life of the community or group. “Work is supposed to be a temporality of slowness. In other words, you don’t rush around like a mad thing in a field, you just quietly and slowly immerse yourself in the field over this extended period of time and get to understand it, get to appreciate it bit by bit.” (minute 7:56).

“Traditional” ethnographers are not necessarily interested in mobility over time or exploring who enters and exits the site. Most ethnographers are only interested in the movement that occurs in the space that is being studied during the time that they are in the field. It is about looking at the roots of the space, not necessarily about looking at the movements into and out of the space.

Multi-sited global ethnography tries to look at the way bounded sites can be studied as unbounded and on the move, as opposed to staying still. It considers how certain things (e.g., things, ideas, people) are  followed as they move. The researcher moves between sites, studying change that is encountered in different sites. From this perspective, the interested lies in the connections between sites. Multiple sites with commonalities can also be studied at the onset, without the need to physically follow.

Paulus, T. M., Lester, J. N., & Dempster, P. G. (2014). Digital Tools for Qualitative Research. Los Angeles, CA: SAGE.

While this text is not solely about ethnographic case studies, it is rich with countless ideas for utilizing digital tools to aid in the multiple facets of qualitative research. In Chapter 5 of their text, entitled Generating Data, the authors dedicate a section to exploring Internet archives and multimedia data. They state that, “in addition to online communities, the Internet is rich with multimedia data such as professionally curated archives, ameteur-created YouTube and Vimeo videos and photo-sharing sites” (p. 81). They provide three specific examples, each explained below: The Internet Archive, CADENSA, and Britain’s BBC Archives.

The Internet Archive ( https://archive.org ) is a non-profit library of millions of free books, movies, software, music, websites, and more. The site also contains a variety of cultural artifacts that are easily available and downloadable. CADENSA ( http://cadensa.bl.uk ) is an online archive of the British Library Sound and Moving Image Catalogue. And finally, the BBC Archives ( http://www.bbc.co.uk/archive/ ) is a particularly useful site for researchers interested in reviewing documentary film and political speeches.

Wang, T. (2016, September). Tricia Wang: The human insights missing from big data. [Video file]. Retrieved from  https://www.ted.com/talks/tricia_wang_the_human_insights_missing_from_big_data

In this TED Talk, Tricia Wang discusses her ethnographic work with technology and advocates for the need to save a place for thick data as opposed to relying only on big data. She argues that while companies invest millions of dollars in generating big data because they assume it will efficiently provide all the answers, it routinely does not provide a good return on investment. Instead, companies are left without answers to the questions about consumer preferences and behaviors, which leaves them unprepared for market changes.

In turn, Wang coins the term thick data, which is described as “precious data from humans, like stories, emotions, and interactions that cannot be quantified” (Minute 11:50). Wang suggests that this thick data may only come from a small group of individuals, but it is an essential component that can provide insights that are different and valuable. As an example, while working for Nokia, her ethnographic experiences in China provided her with new understandings on the future demand for smartphones. However, her employer did not take her findings seriously, and as a result, they lost their foothold in the technology market. She posits that a blended approach to collecting and analyzing data (i.e. combining or integrating thick data analysis with big data analysis) allows for a better grasp on the whole picture and making informed decisions.

Her conclusions for a blended approach to data collection also have implications for blending ethnographic and case-study approaches. While Wang took more of an ethnographic approach to her research, one could envision what her work might have looked like if she had used an Ethnographic Case Study approach. Wang could have clearly defined the time and space boundaries of her various ethnographic experiences (e.g. as a street vendor, living in the slums, hanging out in internet cafés). This would have allowed her to infer causality through the generation of thick data with a small sample size for each location and bound by each group.

Ethnographic Case Studies Copyright © 2019 by Jeannette Armstrong; Laura Boyle; Lindsay Herron; Brandon Locke; and Leslie Smith is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License , except where otherwise noted.

Share This Book

Have a language expert improve your writing

Run a free plagiarism check in 10 minutes, generate accurate citations for free.

  • Knowledge Base

Methodology

  • What Is Ethnography? | Definition, Guide & Examples

What Is Ethnography? | Definition, Guide & Examples

Published on March 13, 2020 by Jack Caulfield . Revised on June 22, 2023.

Ethnography is a type of qualitative research that involves immersing yourself in a particular community or organization to observe their behavior and interactions up close. The word “ethnography” also refers to the written report of the research that the ethnographer produces afterwards.

Ethnography is a flexible research method that allows you to gain a deep understanding of a group’s shared culture, conventions, and social dynamics. However, it also involves some practical and ethical challenges.

Table of contents

What is ethnography used for, different approaches to ethnographic research, gaining access to a community, working with informants, observing the group and taking field notes, writing up an ethnography, other interesting articles.

Ethnographic research originated in the field of anthropology, and it often involved an anthropologist living with an isolated tribal community for an extended period of time in order to understand their culture.

This type of research could sometimes last for years. For example, Colin M. Turnbull lived with the Mbuti people for three years in order to write the classic ethnography The Forest People .

Today, ethnography is a common approach in various social science fields, not just anthropology. It is used not only to study distant or unfamiliar cultures, but also to study specific communities within the researcher’s own society.

For example, ethnographic research (sometimes called participant observation ) has been used to investigate  football fans , call center workers , and police officers .

Advantages of ethnography

The main advantage of ethnography is that it gives the researcher direct access to the culture and practices of a group. It is a useful approach for learning first-hand about the behavior and interactions of people within a particular context.

By becoming immersed in a social environment, you may have access to more authentic information and spontaneously observe dynamics that you could not have found out about simply by asking.

Ethnography is also an open and flexible method. Rather than aiming to verify a general theory or test a hypothesis , it aims to offer a rich narrative account of a specific culture, allowing you to explore many different aspects of the group and setting.

Disadvantages of ethnography

Ethnography is a time-consuming method. In order to embed yourself in the setting and gather enough observations to build up a representative picture, you can expect to spend at least a few weeks, but more likely several months. This long-term immersion can be challenging, and requires careful planning.

Ethnographic research can run the risk of observer bias . Writing an ethnography involves subjective interpretation, and it can be difficult to maintain the necessary distance to analyze a group that you are embedded in.

There are often also ethical considerations to take into account: for example, about how your role is disclosed to members of the group, or about observing and reporting sensitive information.

Should you use ethnography in your research?

If you’re a student who wants to use ethnographic research in your thesis or dissertation , it’s worth asking yourself whether it’s the right approach:

  • Could the information you need be collected in another way (e.g. a survey , interviews)?
  • How difficult will it be to gain access to the community you want to study?
  • How exactly will you conduct your research, and over what timespan?
  • What ethical issues might arise?

If you do decide to do ethnography, it’s generally best to choose a relatively small and easily accessible group, to ensure that the research is feasible within a limited timeframe.

Prevent plagiarism. Run a free check.

There are a few key distinctions in ethnography which help to inform the researcher’s approach: open vs. closed settings, overt vs. covert ethnography, and active vs. passive observation. Each approach has its own advantages and disadvantages.

Open vs. closed settings

The setting of your ethnography—the environment in which you will observe your chosen community in action—may be open or closed.

An open or public setting is one with no formal barriers to entry. For example, you might consider a community of people living in a certain neighborhood, or the fans of a particular baseball team.

  • Gaining initial access to open groups is not too difficult…
  • …but it may be harder to become immersed in a less clearly defined group.

A closed or private setting is harder to access. This may be for example a business, a school, or a cult.

  • A closed group’s boundaries are clearly defined and the ethnographer can become fully immersed in the setting…
  • …but gaining access is tougher; the ethnographer may have to negotiate their way in or acquire some role in the organization.

Overt vs. covert ethnography

Most ethnography is overt . In an overt approach, the ethnographer openly states their intentions and acknowledges their role as a researcher to the members of the group being studied.

  • Overt ethnography is typically preferred for ethical reasons, as participants can provide informed consent…
  • …but people may behave differently with the awareness that they are being studied.

Sometimes ethnography can be covert . This means that the researcher does not tell participants about their research, and comes up with some other pretense for being there.

  • Covert ethnography allows access to environments where the group would not welcome a researcher…
  • …but hiding the researcher’s role can be considered deceptive and thus unethical.

Active vs. passive observation

Different levels of immersion in the community may be appropriate in different contexts. The ethnographer may be a more active or passive participant depending on the demands of their research and the nature of the setting.

An active role involves trying to fully integrate, carrying out tasks and participating in activities like any other member of the community.

  • Active participation may encourage the group to feel more comfortable with the ethnographer’s presence…
  • …but runs the risk of disrupting the regular functioning of the community.

A passive role is one in which the ethnographer stands back from the activities of others, behaving as a more distant observer and not involving themselves in the community’s activities.

  • Passive observation allows more space for careful observation and note-taking…
  • …but group members may behave unnaturally due to feeling they are being observed by an outsider.

While ethnographers usually have a preference, they also have to be flexible about their level of participation. For example, access to the community might depend upon engaging in certain activities, or there might be certain practices in which outsiders cannot participate.

An important consideration for ethnographers is the question of access. The difficulty of gaining access to the setting of a particular ethnography varies greatly:

  • To gain access to the fans of a particular sports team, you might start by simply attending the team’s games and speaking with the fans.
  • To access the employees of a particular business, you might contact the management and ask for permission to perform a study there.
  • Alternatively, you might perform a covert ethnography of a community or organization you are already personally involved in or employed by.

Flexibility is important here too: where it’s impossible to access the desired setting, the ethnographer must consider alternatives that could provide comparable information.

For example, if you had the idea of observing the staff within a particular finance company but could not get permission, you might look into other companies of the same kind as alternatives. Ethnography is a sensitive research method, and it may take multiple attempts to find a feasible approach.

All ethnographies involve the use of informants . These are people involved in the group in question who function as the researcher’s primary points of contact, facilitating access and assisting their understanding of the group.

This might be someone in a high position at an organization allowing you access to their employees, or a member of a community sponsoring your entry into that community and giving advice on how to fit in.

However,  i f you come to rely too much on a single informant, you may be influenced by their perspective on the community, which might be unrepresentative of the group as a whole.

In addition, an informant may not provide the kind of spontaneous information which is most useful to ethnographers, instead trying to show what they believe you want to see. For this reason, it’s good to have a variety of contacts within the group.

Here's why students love Scribbr's proofreading services

Discover proofreading & editing

The core of ethnography is observation of the group from the inside. Field notes are taken to record these observations while immersed in the setting; they form the basis of the final written ethnography. They are usually written by hand, but other solutions such as voice recordings can be useful alternatives.

Field notes record any and all important data: phenomena observed, conversations had, preliminary analysis. For example, if you’re researching how service staff interact with customers, you should write down anything you notice about these interactions—body language, phrases used repeatedly, differences and similarities between staff, customer reactions.

Don’t be afraid to also note down things you notice that fall outside the pre-formulated scope of your research; anything may prove relevant, and it’s better to have extra notes you might discard later than to end up with missing data.

Field notes should be as detailed and clear as possible. It’s important to take time to go over your notes, expand on them with further detail, and keep them organized (including information such as dates and locations).

After observations are concluded, there’s still the task of writing them up into an ethnography. This entails going through the field notes and formulating a convincing account of the behaviors and dynamics observed.

The structure of an ethnography

An ethnography can take many different forms: It may be an article, a thesis, or an entire book, for example.

Ethnographies often do not follow the standard structure of a scientific paper, though like most academic texts, they should have an introduction and conclusion. For example, this paper begins by describing the historical background of the research, then focuses on various themes in turn before concluding.

An ethnography may still use a more traditional structure, however, especially when used in combination with other research methods. For example, this paper follows the standard structure for empirical research: introduction, methods, results, discussion, and conclusion.

The content of an ethnography

The goal of a written ethnography is to provide a rich, authoritative account of the social setting in which you were embedded—to convince the reader that your observations and interpretations are representative of reality.

Ethnography tends to take a less impersonal approach than other research methods. Due to the embedded nature of the work, an ethnography often necessarily involves discussion of your personal experiences and feelings during the research.

Ethnography is not limited to making observations; it also attempts to explain the phenomena observed in a structured, narrative way. For this, you may draw on theory, but also on your direct experience and intuitions, which may well contradict the assumptions that you brought into the research.

If you want to know more about statistics , methodology , or research bias , make sure to check out some of our other articles with explanations and examples.

  • Normal distribution
  • Degrees of freedom
  • Null hypothesis
  • Discourse analysis
  • Control groups
  • Mixed methods research
  • Non-probability sampling
  • Quantitative research
  • Ecological validity

Research bias

  • Rosenthal effect
  • Implicit bias
  • Cognitive bias
  • Selection bias
  • Negativity bias
  • Status quo bias

Cite this Scribbr article

If you want to cite this source, you can copy and paste the citation or click the “Cite this Scribbr article” button to automatically add the citation to our free Citation Generator.

Caulfield, J. (2023, June 22). What Is Ethnography? | Definition, Guide & Examples. Scribbr. Retrieved July 30, 2024, from https://www.scribbr.com/methodology/ethnography/

Is this article helpful?

Jack Caulfield

Jack Caulfield

Other students also liked, what is qualitative research | methods & examples, what is a case study | definition, examples & methods, critical discourse analysis | definition, guide & examples, get unlimited documents corrected.

✔ Free APA citation check included ✔ Unlimited document corrections ✔ Specialized in correcting academic texts

  • - Google Chrome

Intended for healthcare professionals

  • 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
  • Qualitative research...

Qualitative research methodologies: ethnography

  • Related content
  • Peer review
  • Scott Reeves , associate professor 1 ,
  • Ayelet Kuper , assistant professor 2 ,
  • Brian David Hodges , associate professor and vice chair (education) 3
  • 1 Department of Psychiatry, Li Ka Shing Knowledge Institute, Centre for Faculty Development, and Wilson Centre for Research in Education, University of Toronto, 200 Elizabeth Street, Eaton South 1-565, Toronto, ON, Canada M5G 2C4
  • 2 Department of Medicine, Sunnybrook Health Sciences Centre, and Wilson Centre for Research in Education, University of Toronto, Toronto, ON, Canada M4N 3M5
  • 3 Department of Psychiatry, Wilson Centre for Research in Education, University of Toronto, Toronto, ON, Canada M5G 2C4
  • Correspondence to: S Reeves scott.reeves{at}utoronto.ca

The previous articles (there were 2 before this 1) in this series discussed several methodological approaches commonly used by qualitative researchers in the health professions. This article focuses on another important qualitative methodology: ethnography. It provides background for those who will encounter this methodology in their reading rather than instructions for carrying out such research.

What is ethnography?

Ethnography is the study of social interactions, behaviours, and perceptions that occur within groups, teams, organisations, and communities. Its roots can be traced back to anthropological studies of small, rural (and often remote) societies that were undertaken in the early 1900s, when researchers such as Bronislaw Malinowski and Alfred Radcliffe-Brown participated in these societies over long periods and documented their social arrangements and belief systems. This approach was later adopted by members of the Chicago School of Sociology (for example, Everett Hughes, Robert Park, Louis Wirth) and applied to a variety of urban settings in their studies of social life.

The central aim of ethnography is to provide rich, holistic insights into people’s views and actions, as well as the nature (that is, sights, sounds) of the location they inhabit, through the collection of detailed observations and interviews. As Hammersley states, “The task [of ethnographers] is to document the culture, the perspectives and practices, of the people in these settings. The aim is to ‘get inside’ the way each group of people sees the world.” 1 Box 1 outlines the key features of ethnographic research.

Box 1 Key features of ethnographic research 2

A strong emphasis on exploring the nature of a particular social phenomenon, rather than setting out to test hypotheses about it

A tendency to work primarily with “unstructured data” —that is, data that have not been coded at the point of data collection as a closed set of analytical categories

Investigation of a small number of cases (perhaps even just …

Log in using your username and password

BMA Member Log In

If you have a subscription to The BMJ, log in:

  • Need to activate
  • Log in via institution
  • Log in via OpenAthens

Log in through your institution

Subscribe from £184 *.

Subscribe and get access to all BMJ articles, and much more.

* For online subscription

Access this article for 1 day for: £50 / $60/ €56 ( excludes VAT )

You can download a PDF version for your personal record.

Buy this article

ethnographic case study methodology

An Ethnographic Case Study Design

  • First Online: 23 January 2020

Cite this chapter

ethnographic case study methodology

  • Congjun Mu 2  

461 Accesses

This chapter justifies the selection of an ethnographic case study approach to investigate Chinese multilingual scholars’ experiences in writing for scholarly publication in English. Mixed methods—quantitative survey and qualitative semi-structured interviews—are used to elicit data exposing Chinese scholars’ attitude to the controversies discussed in the literature and their strategies to cope with the challenges they face in writing and publishing in English. An in-depth case analysis method with text-history analysis is introduced. The questionnaire design owes much to previous studies in ERPP research, a field that has developed rapidly in recent years and may continue to grow in the future. The procedures of semi-structured interviews and document collection are presented in detail in the interests of transparency, reliability, and validity of the research.

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

Access this chapter

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
  • Durable hardcover edition

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Bazerman, C., Keranen, N., & Encinas, F. (2012). Facilitated immersion at a distance in second language scientific writing. In M. C. Badia & C. Donahue (Eds.), University Writing: Selves and Texts in Academic Societies (pp. 235–252). London: Emerald.

Google Scholar  

Burgess, S., Gea-Valor, M. L., Moreno, A. I., & Rey-Rocha, J. (2014). Affordances and constraints on research publication: A comparative study of the language choices of Spanish historians and psychologists. Journal of English for Academic Purposes, 14 , 72–83.

Article   Google Scholar  

Casanave, C. P. (1998). Transitions: The balancing act of bilingual academics. Journal of Second Language Writing, 7 (2), 17–203.

Cho, D. W. (2009). Science journal paper writing in an EFL context: The case of Korea. English for Specific Purposes, 28 (4), 230–239.

Counsell, J. (2011). How effectively and consistently do international postgraduate students apply the writing strategies they have been taught in a generic skills-based course to their subsequent discipline-based studies? Journal of Academic Language & Learning, 5 (1), A1–A17.

Curry, M. J., & Lillis, T. (2004). Multilingual scholars and the imperative to publish in English: Negotiating interests, demands, and rewards. TESOL Quarterly, 38 (4), 663–688.

Curry, M. J., & Lillis, T. M. (2010). Academic research networks: Accessing resources for English-medium publishing. English for Specific Purposes, 29 , 281–295.

Fernández Polo, F. J., & Cal Varela, M. (2009). English for research purposes at the University of Santiago de Compostela: A survey. Journal of English for Academic Purposes, 8 , 152–164.

Fife, W. (2005). Doing Fieldwork Ethnographic Methods for Research in Developing Countries and Beyond . London: Palgrave Macmillan.

Flowerdew, J. (1999a). Problems in writing for scholarly publication in English: The case of Hong Kong. Journal of Second Language Writing, 8 , 243–263.

Flowerdew, J. (1999b). Writing for scholarly publication in English: The case of Hong Kong. Journal of Second Language Writing, 8 , 123–145.

Flowerdew, J. (2000). Discourse community, legitimate peripheral participation, and the nonnative-English-speaking scholar. TESOL Quarterly, 34 (1), 127–150.

Flowerdew, J. (2008). Scholarly writers who use English as an Additional Language: What can Goffman’s “Stigma” tell us? Journal of English for Academic Purposes, 7 , 77–86.

Flowerdew, J. (2009). Goffman’s stigma and EAL writers: The author responds to Casanave. Journal of English for Academic Purposes, 8 , 69–72.

Fusch, P. I., Fusch, G. E., & Ness, L. R. (2017). How to conduct a mini-ethnographic case study: A guide for novice researchers. The Qualitative Report, 22 (3), 923–941.

Goetz, J. P., & LeCompte, M. D. (1984). Ethnography and Qualitative Design in Educational Research . London: Academic Press.

Gollin-Kies, S. (2014). Methods reported in ESP research articles: A comparative survey of two leading journals. English for Specific Purposes, 36 , 27–34.

Hanauer, D. I., & Englander, K. (2011). Quantifying the burden of writing research articles in a second language: Data from Mexican scientists. Written Communication, 28 , 403–416.

Hanauer, D. I., & Englander, K. (2013). Scientific Writing in a Second Language . Anderson, SC: Parlor Press.

Hirano, E. (2009). Research article introductions in English for specific purposes: A comparison between Brazilian Portuguese and English. English for Specific Purposes, 28 (4), 240–250.

Holloway, I., Brown, L., & Shipway, R. (2010). Meaning not measurement: Using ethnography to bring a deeper understanding to the participant experience of festivals and events. International Journal of Event and Festival Management, 1 (1), 74–85.

Kourilova, M. (1998). Communicative characteristics of reviews of scientific papers written by non-native users of English. Endocrine Regulation, 32 , 107–114.

LeCompte, M. D., & Goetz, J. P. (1982). Problems of reliability and validity in ethnographic research. Review of Educational Research, 52 (1), 31–60.

LeCompte, M. D., Preissle, J., & Tesch, R. (2008). Ethnography and Qualitative Design in Educational Research (2nd ed.). Bingley, UK: Emerald.

Lei, J., & Hu, G. (2019). Doctoral candidates’ dual role as student and expert scholarly writer: An activity theory perspective. English for Specific Purposes, 54 , 62–74.

Li, X., Xue, M., Wang, Q., & Feng, H. (2011). The thinking and practice on performance evaluation of the classification of higher education institutions in Shanghai. Research in Education Development, 17 , 1–5.

Li, Y. (2002). Writing for international publication: The perception of Chinese doctoral researchers. Asian Journal of English Language Teaching, 12 , 179–194.

Li, Y. (2006a). A doctoral student of physics writing for publication: A sociopolitically-oriented case study. English for Specific Purposes, 25 , 456–478.

Li, Y. (2006b). Negotiating knowledge contribution to multiple discourse communities: A doctoral student of computer science writing for publication. Journal of Second Language Writing, 15 , 159–178.

Li, Y. (2012). “I have no time to find out where the sentences came from; I just rebuild them”: A biochemistry professor eliminating novices’ textual borrowing. Journal of Second Language Writing, 21 , 59–70.

Lillis, T. (2008). Ethnography as method, methodology, and “Deep Theorizing”: Closing the gap between text and context in academic writing research. Written Communication, 25 (3), 353–388.

Lillis, T., & Curry, M. J. (2006). Professional academic writing by multilingual scholars interactions with literacy brokers in the production of English-medium texts. Written Communication, 23 , 3–35.

Marshall, C., & Rossman, G. B. (1999). Designing Qualitative Research . London: SAGE Publications.

Miles, M., & Huberman, A. (1994). Qualitative Data Analysis: An Expanded Sourcebook (2nd ed.). Thousand Oaks, CA: Sage.

Moreno, A. I., Burgess, S., Sachdev, I., López-Navarro, I., & Rey-Rocha, J. (2013). The ENEIDA questionnaire: Publication experiences in scientific journals in English and Spanish. Retrieved from http://eneida.unileon.es/eneidaquestionnaire.php

Moreno, A. I., Rey-Rocha, J., Burgess, S., López-Navarro, I., & Sachdev, I. (2012). Spanish researchers’ perceived difficulty writing research articles for English medium journals: The impact of proficiency in English versus publication experience. Ibérica, 24 , 157–184.

Mur Dueñas, P. (2012). Getting research published internationally in English: An ethnographic account of a team of Finance Spanish scholars’ struggles. Iberica, 24 , 139–156.

Muresan, L.-M., & Pérez-Llantada, C. (2014). English for research publication and dissemination in bi-/multiliterate environments: The case of Romanian academics. Journal of English for Academic Purposes, 13 , 53–64.

O’Malley, J. M., & Chamot, A. U. (1990). Learning Strategies in Second Language Acquisition . New York: Cambridge University Press.

Book   Google Scholar  

Patton, M. Q. (2002). Qualitative Research & Evaluation Methods (3rd ed.). London: Sage.

Ramanathan, V., & Atkinson, D. (1999). Ethnograph approaches and methods in L2 writing research: A critical guide and review. Applied linguistics, 20 (1), 41–70.

Stake, R. E. (2010). Qualitative Research: Studying How Things Work . New York: Guilford Press.

Teng, L. S., & Zhang, L. J. (2016). A questionnaire-based validation of multidimensional models of self-regulated learning strategies. The Modern Language Journal, 100 (3), 674–701.

Watson-Gegeo, K. A. (1988). Ethnography in ESL: Defining the essentials. TESOL Quarterly, 22 (4), 575–592.

Yanow, D., & Schwartz-Shea, P. (Eds.). (2006). Interpretation and Method Empirical Research Methods and the Interpretive Turn . New York: M. E. Sharpe.

Yin, R. K. (2014). Case Study Research: Designs and Methods (5th ed.). Thousand Oaks, CA: Sage.

Yockey, R. D. (2010). SPSS Demystified (C. Liu & Z. Wu, Chinese Trans.). Beijing: Renmin University of China Press.

Zhu, W. (2004). Faculty views on the importance of writing, the nature of academic writing, and teaching and responding to writing in the disciplines. Journal of Second Language Writing, 13 , 29–48.

Download references

Author information

Authors and affiliations.

College of Foreign Languages, Shanghai Maritime University, Shanghai, China

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Congjun Mu .

Rights and permissions

Reprints and permissions

Copyright information

© 2020 The Author(s)

About this chapter

Mu, C. (2020). An Ethnographic Case Study Design. In: Understanding Chinese Multilingual Scholars’ Experiences of Writing and Publishing in English. Palgrave Macmillan, Cham. https://doi.org/10.1007/978-3-030-33938-8_4

Download citation

DOI : https://doi.org/10.1007/978-3-030-33938-8_4

Published : 23 January 2020

Publisher Name : Palgrave Macmillan, Cham

Print ISBN : 978-3-030-33937-1

Online ISBN : 978-3-030-33938-8

eBook Packages : Social Sciences Social Sciences (R0)

Share this chapter

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

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research
  • Open access
  • Published: 05 December 2021

Ethnographic research as an evolving method for supporting healthcare improvement skills: a scoping review

  • Georgia B. Black   ORCID: orcid.org/0000-0003-2676-5071 1 ,
  • Sandra van Os   ORCID: orcid.org/0000-0003-0021-8758 1 ,
  • Samantha Machen   ORCID: orcid.org/0000-0003-4727-4423 1 &
  • Naomi J. Fulop   ORCID: orcid.org/0000-0001-5306-6140 1  

BMC Medical Research Methodology volume  21 , Article number:  274 ( 2021 ) Cite this article

23k Accesses

6 Citations

14 Altmetric

Metrics details

A Correction to this article was published on 11 April 2022

This article has been updated

The relationship between ethnography and healthcare improvement has been the subject of methodological concern. We conducted a scoping review of ethnographic literature on healthcare improvement topics, with two aims: (1) to describe current ethnographic methods and practices in healthcare improvement research and (2) to consider how these may affect habit and skill formation in the service of healthcare improvement.

We used a scoping review methodology drawing on Arksey and O’Malley’s methods and more recent guidance. We systematically searched electronic databases including Medline, PsychINFO, EMBASE and CINAHL for papers published between April 2013 – April 2018, with an update in September 2019. Information about study aims, methodology and recommendations for improvement were extracted. We used a theoretical framework outlining the habits and skills required for healthcare improvement to consider how ethnographic research may foster improvement skills.

We included 274 studies covering a wide range of healthcare topics and methods. Ethnography was commonly used for healthcare improvement research about vulnerable populations, e.g. elderly, psychiatry. Focussed ethnography was a prominent method, using a rapid feedback loop into improvement through focus and insider status. Ethnographic approaches such as the use of theory and focus on every day practices can foster improvement skills and habits such as creativity, learning and systems thinking.

Conclusions

We have identified that a variety of ethnographic approaches can be relevant to improvement. The skills and habits we identified may help ethnographers reflect on their approaches in planning healthcare improvement studies and guide peer-review in this field. An important area of future research will be to understand how ethnographic findings are received by decision-makers.

Peer Review reports

Research can help to support the practice of healthcare improvement, and identify ways to “improve improvement” [ 1 ]. Ethnography has been identified particularly as a research method that can show what happens routinely in healthcare, and reveal the ‘ what and how of improving patient care [ 2 ]. Ethnography is not one method, but a paradigm of mainly qualitative research involving direct observations of people and places, producing a written account of natural or everyday behaviours and ideas [ 3 ]. Ethnographic research can identify contextual barriers to healthcare improvement. For example, Waring and colleagues suggested that hospital discharge could be improved by allowing staff to have more opportunities for informal communication [ 4 ].

There have been advances in ethnographic methods that support its role in supporting healthcare improvement. Multi-site, collaborative modalities of ethnography have evolved that suit the networked nature of modern healthcare [ 5 ]. Similarly, rapid ethnographic approaches (e.g. Bentley et al. [ 6 ];) meet the needs of improvement activities to produce findings within short timeframes [ 7 ]. However, the production of sustained ethnographic fieldwork has waned in response to demands for rapid evidence [ 6 , 8 , 9 ]. Critics of rapid ethnographic methods worry that they are diluting ethnography within applied contexts more widely [ 5 , 10 ].

The relationship between ethnography and healthcare improvement has been the subject of methodological concern [ 8 ]. The first concern is that some research identified as ethnography does not fit within the ethnographic paradigm, merely collecting observational data without a theoretical analysis, interpretation or researcher reflexivity [ 11 ]. A second concern is whether the topics of ethnographic inquiry produce findings that are seen as useful for improvement [ 12 ], particularly if they do not make explicit recommendations or produce checklists [ 8 , 13 , 14 , 15 ]. Authors fear that ethnographic findings that capture complexity [ 16 ] and expose taken-for-granted behaviours and phenomena [ 14 , 17 ] may be too abstract to be relevant to healthcare improvement [ 8 ]. However, these critiques position ethnographic research as a product which may be taken up by healthcare improvers, rather than seeing ethnographic work itself as an improvement activity. We take the view that healthcare improvement aims to change human behaviour to improve patient care, and is therefore reliant on the development of particular skills and habits (such as good communication) [ 18 ]. We would consider that engaging in ethnographic research may support skill development and habit formation that serves healthcare improvement.

In the literature of ethnography in healthcare improvement, there is not much discussion of the close relationship between methodological features of ethnographic research, and their impact on improvement skills. The aim of this paper is twofold: (1) to describe current ethnographic methods and practices in healthcare improvement research and (2) to consider how these may affect habit and skill formation in the service of healthcare improvement [ 19 ].

This is a scoping review following the methods outlined by Arksey & O’Malley and later refined by Levac et al., [ 20 , 21 ] including a systematically conducted literature review and reported in accordance with the Preferred Reporting Items for Systematic reviews and Meta-Analyses extension for Scoping Reviews (PRISMA-ScR; see Additional file 1 for PRISMA checklist). No protocol was published for this review. Our literature search and analyses were conducted iteratively, searching reference lists and undertaking discussions with colleagues about key lines of argument. We also held a workshop at Health Services Research UK conference in 2018 on this topic to gain a wide range of stakeholder views.

Systematic retrieval of empirical papers and purposive sampling

Our search strategy was designed to capture a wide range of approaches to ethnography from different journals, healthcare settings and types of research environment. It was not our aim to capture every study using this methodology, but to map the current field. Thus we did not search grey literature, books or monographs. The search strategy was developed and piloted in consultation with a health librarian. Medline (on OVID platform), PsychINFO, CINAHL and EMBASE databases were searched, and six journals were hand-searched, including: BMJ Quality & Safety, Social Science and Medicine, Medical Anthropology, Cochrane library, Sociology of Health and Illness and Implementation Science. These databases were searched between dates April 2013 – April 2018 and an update was performed in September 2019 using the search terms outlined in Additional file 2 . We limited the search to these dates in order to capture the most recent methodological characteristics of ethnographic studies in this field.

We screened titles and then abstracts according to the inclusion and exclusion criteria detailed in Table 1 . We included studies which self-identified as using ethnography or ethnographic methods rather than using our own criteria. This is because ethnography can be hard to define, and use of criteria may risk excluding papers which exemplify the sorts of tensions and workarounds we are trying to capture.

The retrieved papers were screened by GB, SVO and SM based on inclusion and exclusion criteria (Table 1 ). The total number of papers after screening titles, abstracts and full texts was 274 (Fig. 1 ).

figure 1

PRISMA statement of all references retrieved, screened and included in the scoping review

Numerical charting

Characteristics of each paper, such as title, authors, journal, year, country and healthcare subject area were extracted (see Table 2 ).

Thematic analysis and development

We coded all 274 papers using NVivo software for stated aims and recommendations. This included close reading, and retrieval of key ideas and quotations from the papers that exemplified key ideas in relation to healthcare improvement, methodology and the authors’ reflections on these. The coded extracts of aims and recommendation in conjunction with the closer reading of the sub-sample were used to inductively develop conceptual ideas, such as how the corpus of papers explicitly aimed to contribute to healthcare improvement, and if not, how this affected the types of conclusions drawn. Some papers were read in greater depth to understand how the authors’ methods related to their findings and conclusions. In order to consider how ethnography supports habits and skills associated with healthcare improvement, we drew on a framework which identifies five habits of ‘improvers’: creativity, learning, systems thinking, resilience and influencing [ 19 ]. Applying this model to our selected papers, we mapped traits or approaches to the ethnographic studies that exemplified these habits either in the authors, or as part of developing these habits in others (e.g. healthcare decision-makers and professionals). Thematic interpretations and lines of argument were generated and discussed by all the authors.

Overview of study characteristics

The included studies covered a wide range of ethnographic methodologies and healthcare subjects, published internationally (Table 2 ) in predominantly social science and clinical journals (see Additional file 3 ). The full list of the 274 included studies is available in Additional file 4 .

Most studies described themselves as an ‘ethnography’ or ‘ethnographic’, although some described their methodology as ‘mixed methods’ including ethnographic components. For example, Collet et al. conducted a mixed methods participatory action research study using observations to produce an “ethnographic description” [ 22 ].

Almost all studies relied on observation and interviews as the main data sources. It was not always specified whether researchers took a participant or non-participant approach to observation. There were some examples of other data sources e.g. video data, surveys, documents, field notes, diaries, and artefacts. A few examples contained a paucity of data, such as only video data [ 23 ], limited fieldwork [ 24 ], a small number of interviewees [ 25 ], or reliance on focus group data alone [ 26 ]. Methods associated with qualitative methodology (but not necessarily ethnographic) were also used, such as data ‘saturation’ to denote that additional data did not provide new insights into the topic [ 27 ].

There were a number of minority or unusual ethnographic variations:

Quantitative ethnography [ 23 ]: temporal coding of physicians' workflow and interaction with the electronic health record system, and their patient.

Cognitive ethnography [ 28 ]: “identifying and elaborating distributed cognitive processes that occur when an individual enacts purposeful improvements in a clinical context”.

Street-level organizational ethnography [ 29 ]: intensive case study methods to explore the implications of healthcare policy at a street level.

Phenomenological ethnographies [ 30 ]: focussing on the lived experience and meanings associated with a phenomenon.

Geo-mapping [ 31 ]: geomapping of selected service data to define Latino immigrant community before conducting interviews and observations.

Use of different types of ethnography to support healthcare improvement

We found that many studies used methods that could identify issues relating to power and vulnerability, with potential relevance to how healthcare improvement problems are defined and solved, and by whom [ 1 ]. For example we noted a significant minority of studies using institutional and critical ethnography, mostly in vulnerable populations (see Table 3 ). These studies were explicitly attentive to systems and power relations, rather than on individual practices. We suggest that the use of geographically-oriented methods such as geo-mapping and street-level organisational ethnography are also attentive to the power structures inherent in place and space, and could be relevant to other geographical healthcare improvement topics such as networked healthcare systems, care at home and patient travel for treatment.

The high prevalence of ethnographic studies with vulnerable populations (e.g. psychiatry, end of life care) suggests that ethnography is also being conceptualised as an emancipatory method, reversing healthcare power structures in its focus. This has been a traditional focus of ethnography since social changes in power and representation in the 1970s, incorporated into the development of healthcare research methodology [ 40 , 41 ]. Some methods used were calculated to maximise the potential for supporting vulnerable groups, for example, Nightingale et al. [ 42 ] used focused ethnography (prolonged fieldwork in a small number of settings) to look at patient-professional interactions in paediatric chronic illness settings. The authors suggested that focussed ethnography is particularly suited to settings where fostering trust is essential. We would also suggest that ethnography may be particularly suited to settings in which participants are less able to verbalise their experiences.

The reviewed studies suggested that video ethnography can support healthcare improvement at a team level. For example, Stevens et al. [ 43 ] promoted video ethnography as a way to capture in-depth data on intimate interactions, in their study of elective caesareans. The video data allowed them to make use of timing data (e.g. of certain actions), physical positioning of different actors and equipment, and verbatim dialogue recording. The video data also suited the technical nature of the procedure, which was relatively time-limited. This form of data collection may not suit environments where healthcare activities are more spread out.

The impact of healthcare practitioner involvement in ethnographic fieldwork and findings

We noted that the use of ethnography for healthcare improvement has led to healthcare practitioners’ widespread involvement in data collection or analysis. We suggest that this is a form of negotiation across the healthcare-academia boundary, translating from ‘real world’ to data and back again. This has potential to create rich and relevant ethnographic studies that are geared towards improvement. However, some studies were undermined by a lack of reflexivity about the dual practitioner-ethnographer role.

A significant number of papers involved healthcare practitioners in fieldwork (e.g. Abdulrehman, 2017, Hoare et al. 2013; [ 37 , 44 ]). For example in Hoare et al. the lead researcher was a nurse, and wrote that they hoped “to bring both an emic and etic perspective to the data collection by bracketing my emic sense of self as a nurse practitioner in order to become a participant observer within my own general practice ” [ 37 ]. In this study, the findings fed directly into local service improvement as the lead researcher felt compelled to “share new ‘best practice’ information and join in the conversation.” There was little discussion about how this affected the generalisability of the findings, and whether their recommendations were adopted.

Similarly, Bergenholz et al. [ 45 ] conducted a study where a nursing researcher completed the main fieldwork and “assisted the nurses with practical care .” They acknowledged that “This may have caused limitations with regards to ‘blind spots’ in the nursing practice, but that it also gave access to a field that might be difficult for ‘outside-outsiders’ to gain .” However, there was no commentary on where the blind spots or extra access occurred, and how this may have affected the relevance and dissemination of their findings.

How might ethnography support healthcare improvement habits?

In this section, we evaluate the studies included in the review in terms of how their methods relate to improvement. We draw on the idea that successful improvement is based on a set of habits and their related skills acquired through experience and practice [ 19 ]. This section is structured around Lucas’s five habits of ‘improvers’: creativity, learning, systems thinking, resilience and influencing [ 19 ]. Under those headings, we describe the mechanisms by which ethnographic studies can support healthcare improvement habits, using illustrative examples.

Resilience is defined as being adaptable, particularly tolerating calculated risks and uncertainty, and proceeding with optimism. Being able to recover from adverse events is core to improvement, reframing them as opportunities. Adaptation and the ability to bounce back from adverse events and variation are core to improvement.

Tolerating the uncertainty of ethnographic data collection

While we did not relate these traits to any particular ethnographic approach in our studies, we would consider that undertaking any ethnographic project requires resilience, as data collection is inherently exploratory and uncertain. For example, Belanger et al. wanted to know how health care providers and their patients approach patient participation in palliative care decisions. The authors explicitly eschewed the pull to create guidelines or other formalised knowledge, but aimed to explore the “unforeseen and somewhat unavoidable ways in which discursive practices prompt or impede patient participation during these interactions.” [ 46 ]

Creativity is defined as working together to encourage fresh thinking by generating ideas and thinking critically.

Using a theoretical lens

Researchers may consider healthcare through a particular theory or framework (e.g. private ordering [ 47 ], masculine discourse [ 48 ], compassion [ 49 ]). The restriction of the theoretical lens enables critical thinking, and keeps the ethnographer creatively engaged. For example, Mylopoulos & Farhat [ 28 ] used the concept of adaptive expertise in a cognitive ethnography to explore “the phenomenon of purposeful improvement” in a teaching hospital. This theoretical lens revealed that clinicians were engaging in “invisible” improvement in their daily work, in “specific activities such as scheduling, establishing patient relationships, designing physical space and building supporting resources”. The authors suggested that these practices were devalued in comparison to more formal improvement activities, justifying the utility of the ‘adaptive expertise’ theory in bringing the daily improvement practices to light.

Challenging current problems and perspectives

We identified studies that challenged or reframed existing improvement problems e.g. Mishra [ 50 ]. This role removes the ‘blinkers’ of improvement research [ 51 ], and can ‘dissolve’ previously intractable implementation problems. For example, Boonan et al. [ 52 ] studied the practice of bar-coded medication from the perspective of nurses using the intervention. In their discussion, the authors challenge the assumption that if you introduce technology, then you will mitigate human factor risks. They highlighted that external pressures on hospitals perpetuate this perspective, and that “nurses and patients are consequently drawn into this discourse and institutional ruling, to which they are not oblivious”. Their recommendation was to understand the skills of nurses in tailoring technology to meet individual patients’ needs rather than trusting in systems blindly.

Learning is defined as harnessing curiosity and using reflective processes to extract meaning from experience.

Inviting reflection

We noted that some studies did not make explicit recommendations for improvement, but wrote their findings in a manner that would invite reflection on its subject matter. For example, Thomas & Latimer [ 53 ] wrote that they view their role as provocateurs of new ideas, stating that their intention “is not to propose specific policies or discourses designed to change or improve practice. More modestly, we hope that by analysing the everyday and by theorising the mundane, this article will ignite reflexive, ethical and pluralistic dialogues – and so better communication between practitioners, parents and the wider lay public – around reproductive technologies and medical conditions” (authors’ underline; p.951-2) [ 53 ]. Others such as Mackintosh et al [ 54 ] used their discussion section to examine their results in the context of other theories and provide illumination: “Our focus on trajectories illuminates the physiological process of birth and the unfolding pathology of illness (and death). This frame provides a means for us to link the agency of those involved in organising the care of acutely ill patients with the wider socio-political factors beyond the clinic, such as governmentality and risk (Heyman 2010, Waring 2007), death brokering (Timmermans 2005) and the medicalisation of birth and death (De Vries 1981).” (p.264). These two examples show that ethnographic work can be offered as an opportunity for learning and reflection, without a translation to specific recommendations.

Supporting a more ethical, expansive, inclusive, and participatory mode of healthcare

Problem-finding is highlighted as an important part of learning in improvement [ 19 ]. Several studies paid attention to multivocality and power, using this to find problematic, unethical and exclusive practices in healthcare. For example, some studies reported previously unheard viewpoints [ 55 , 56 , 57 ], or identified restrictive organisational barriers and normative assumptions [ 58 , 59 ]. Others promoted ethnography as a way of exploring ethics and morality [ 47 , 60 , 61 ], such as criticising research that prioritizes the needs of individuals over the good of society [ 62 ]. Ross et al. [ 63 ] suggested that it is also more ethical to use critical ethnography than other evaluative methods in researching vulnerable populations (e.g. neurological illness), by being able to “explore perceived political and emancipatory implications, [clarify] existing power differentials and [maintain] an explicit focus on action” .

Some studies directly researched power within the healthcare setting. For example, Batch and Windsor’s study of nursing workforce suggested that senior nurse leaders should use their positions to advocate for better working conditions [ 35 ], “ Manageable nurse/patient ratios, flexible patient-centred work models, equal opportunity for advancement, skill development for all and unit teamwork promotion”. Challenging traditional cultural assumptions that have produced and reproduced stereotypes is problematic because they most often are, by their very nature, invisible. In a more critical approach, Gesbeck’s thesis [ 62 ] on diabetes care work challenges the very mechanism of achieving healthcare improvement through research, stating that “we need to change the social and political context in which health care policy is made. This requires social change that prioritizes the good of the society over the good of the individual—a position directly opposed to the current system oriented toward profit and steeped in the ideology of personal responsibility.”

Systems thinking

Systems thinking is defined as seeing whole systems as well as their parts and recognising complex relationships, connections and interdependencies.

Suggesting reorientation to new ‘problem’ areas

We found that many ethnographic studies emphasised skills of synthesis and connection-making, reorienting improvement to different areas, for example in overarching policy recommendations (e.g. Hughes [ 36 ]; Liu et al. [ 64 ], Matinga et al. [ 65 ]), or resetting priorities. For example, Manias’ [ 66 ] ethnography of communication relating to family members' involvement in medication management in hospital suggests that “greater attention should be played on health professionals initiating communication in proactive ways ” [p.865]. In another example, Cable-Williams & Wilson’s (2017) focussed ethnography captures cultural factors within long-term care facilities. Their discussion suggests that acknowledgement of death is under-represented in front-line practice and government policy, reorienting discussions towards an integration of living and dying care.

Exposing hidden practices within the everyday

We found that several studies drew attention to ‘hidden’ practices in healthcare work, allowing them to evaluated and improved. For example, we found reference to practices such as coordinating [ 67 ], repair [ 68 ], caretaking [ 69 ], scaffolding [ 68 ], tinkering [ 52 ] and bricolage [ 58 ]. We also found that some studies had new interpretations of ‘the everyday’ or ‘taken-for-granted’ (e.g. nursing culture [ 34 , 35 , 45 , 70 ], interprofessional practice [ 67 , 71 , 72 , 73 , 74 , 75 ]). Authors’ outputs included frameworks [ 76 ] or models [ 69 , 71 , 77 , 78 ] that map these types of practices in a way that is helpful for intervention development or quality improvement. For example, Mackintosh et al. [ 54 ] looked at rescue practices in medical wards and maternity care settings using Strauss’s concept of the patient trajectory. Their findings highlighted the risks inherent in the wider social practices of hospital care, and suggested that improvement was needed at a level “beyond individual and team processes and technical safety solutions.”

Influencing

Influencing is defined as engaging others and gaining buy-in using a range of facilitative processes.

Direct translation of findings to targets for improvement

Lucas suggests that to be influential, ethnographic studies need to have some empathy with clinical reality, whilst being facilitative and comfortable with conflict [ 19 ]. This was shown in ethnographic studies that made pragmatic recommendations, such as in Jensen’s study of clinical simulation. They advised that simulation might be useful in staging “adverse event scenarios with a view to creating more controlled and safer environments.” ( 80). In MacKichan et al. [ 79 ] observations and interviews were used to understand how primary care access influenced decisions to seek help at the emergency department. The authors made empathic, actionable recommendations such as “ simplifying appointments systems and communicating mechanisms to patients.” (p.10).

Evaluating the context of healthcare improvement

By capturing contextual and social aspects of healthcare improvement, ethnographic evaluations can support leaders and managers who are trying to implement improvement activities. This is a particularly helpful trait in ethnographic studies that pay attention to politics, governance and social theory in their evaluation of new interventions, “zooming out” [ 80 ] beyond the patient-clinician interaction to broader social networks. For example, Tietbohl et al. [ 81 ] investigated the difficulties of implementing a patient decision support intervention (DESI) in primary care through the theoretical lens of relational coordination between “physician and clinical staff groups (healthcare professionals)”. The authors’ recommended attention to the “underlying barriers such as the relational dynamics in a medical clinic or healthcare organization” when creating policies and programs that support shared decision-making using support interventions. This sort of insight can make it more likely that new policies or interventions will succeed. This skill was particularly fertile in the tradition of techno-anthropology, exploring technology-induced errors and the real-world interaction between people and technology, e.g. decision-support tools [ 81 , 82 , 83 , 84 , 85 , 86 ], the introduction of robot caregivers [ 87 ] and clinical simulations [ 88 ]. Other approaches included an investigation of one intervention or change but with a theoretical lens of inquiry.

Summary of findings

This scoping review has identified the methodological characteristics of 5 years of published papers that self-identify as ethnography or ethnographic in the field of healthcare improvement. Ethnography is currently a popular research method in a wide range of healthcare topics, particularly in psychiatry, e.g. mental health, dementia and experiential concerns such as quality of life. Focused ethnography is a significant sub-group in healthcare, suggesting that messages about the importance of research timeliness have taken hold [ 89 ].

We have identified ethnographic methods reported in these papers, and considered their utility in developing skills and habits that support healthcare improvement. Specific practices associated with the ethnographic paradigm can encourage good habits (resilience, creativity, learning, systems thinking and influencing) in healthcare, which can support improvement. For example, using relevant theories to look at every day work in healthcare can foster creativity. The use of critical and institutional ethnography could increase skills in ‘systems thinking’ by critically evaluating how healthcare improvement problems are defined and solved, and by whom.

Comparison with previous literature

This scoping review is the first to consider how current ethnographic methods and practices may relate to healthcare improvement. Within the paradigm of applied healthcare research, there is normative value in being ‘useful’ or ‘impactful’ in our research, which affects our prospects for funding and career success [ 12 ]. However, our review has uncovered a multitude of ways that an ethnographic study can be useful in relation to healthcare improvement, without creating actionable findings. We found a spectrum of interactions with healthcare improvement: some authors explicitly eschewed recommendations or clinical implications; others made imperative statements about required changes to policy or practice. However, this diversity was not necessarily a reflection on how ‘traditional’ the ethnographic methodology was. This challenges the paper by Leslie et al. which puts ethnographic studies in two output categories with respect to healthcare improvement: critique versus feedback [ 8 ]. Instead, we uncovered a variety of ways that ethnography can support healthcare improvement habits, such as encouraging reflection, problem-finding and exposing hidden practices in healthcare.

We did find that supporting healthcare improvement through ethnographic research can require strategic effort, however. For example, we noted that several authors wrote multiple articles based on the same project, often for different types of journal to reach different audiences such as diverse readerships in health services and academic settings. For example, Collier and colleagues published two papers based on a video ethnography of end-of-life care (both in 2016), one in a healthcare quality journal [ 32 ] and one in a qualitative research journal [ 76 ]. The former is shorter, with explicit recommendations for patient safety, whereas the latter is longer, has more detailed results and long sections on reflexivity. Similarly, Grant published an article in a sociology journal [ 90 ] and a healthcare improvement paper [ 91 ] on the same work about medication safety. The sociological paper covered “spatio-temporal elements of articulation work” whereas the other put forward “key stages” and risks, suggesting that it was more closely oriented to improvement.

There have been some considerable debates about changes in ethnographic methods and tools, with concerns about lost researcher identity, dilution of the method, and challenges to “upholding ethnographic integrity” [ 92 ] . We contest this, suggesting that new variants such as focussed and cognitive ethnography are evolving in response to the complexity of hospitals and healthcare [ 93 ], while also being highly regulated, standardised and ordered by biomedicine. Such complex environments cannot be studied and improved under one paradigm alone. Ethnographic identity and method have also been affected by the cross-pollination of ethnography with other social science paradigms and applied environments (e.g. clinical trials, technology development). Debates about theoretical and methodological choices are not only made merely with respect to healthcare improvement, but also in response to professional pressures (e.g. university requirements for impact) [ 12 ], and the mores of taste situated within the overlapping communities of practice that evaluate ethnographic healthcare research [ 94 ]. That said, we echo previous authors’ calls for attention to reflexivity, particularly in embedded or clinician-as-researcher roles [ 95 ].

Our scoping review challenges a previously expressed concern that ethnographic studies may not produce findings that are useful for improvement [ 10 , 12 , 16 ]. By considering different ethnographic designs in relation to skills and habits needed for improvement, we have shown that studies need not necessarily produce ‘actionable findings’ in order to make a valuable contribution. Instead, we would characterise ethnography’s role in the canon of healthcare research methodologies as a way of enhancing improvement habits such as comfort with conflict, problem-finding and connection-making.

Strengths and limitations

This review has a number of limitations. The search may not have found all relevant studies, however the retrieved papers are intended as an exemplar rather than an exhaustive or aggregative review. The review is also limited to journal articles as evidence of researchers’ approach to improvement. This ignores many other ‘offline’ and ‘online’ activities such as meetings, presentations, blogs, books, and websites, which are conducted to disseminate findings and ideas. Our reliance on self-report for the identification of ethnographic studies will have excluded some studies within an ethnographic paradigm who chose different terms for their methodology (e.g. critical inquiry, case study). The strengths of this paper are its comprehensive coverage, incorporating all representative studies in healthcare research published within a five year period, and a wide range of ethnographic sub-types and healthcare subjects, drawn from an international pool of research communities.

We did not prescribe the right way for ethnographers to engage in healthcare improvement, indeed, we have identified that a variety of approaches can be relevant to improvement. The habits we identified may help ethnographers reflect on their approaches in planning healthcare improvement studies and guide peer-review in this field. Issues of taste, traditionalism and researcher identity need to be scrutinised in favour of value and audience. An important area of future research will be to understand how ethnographic findings are received by decision-makers, and further focused reviews on the relationship(s) between ethnographic methods, quality improvement skills and improvement outcomes.

Availability of data and materials

All papers included in the review are listed in Additional file 4 and are publicly available from their publishers’ websites.

Change history

11 april 2022.

A Correction to this paper has been published: https://doi.org/10.1186/s12874-022-01587-9

Dixon-Woods M. How to improve healthcare improvement—an essay by Mary Dixon-Woods. BMJ. 2019;367:l5514.

Article   PubMed   PubMed Central   Google Scholar  

Dixon-Woods M. What can ethnography do for quality and safety in health care? Qual Saf Health Care. 2003;12(5):326.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Savage J. Ethnography and health care. BMJ. 2000;321(7273):1400.

Waring J, Marshall F, Bishop S. Understanding the occupational and organizational boundaries to safe hospital discharge. J Health Serv Res Policy. 2014;20(1_suppl):35–44.

Article   Google Scholar  

Marcus GE. Multi-sited ethnography: Five or six things I know about it now. In: Coleman S, von Hellerman P, editors. Multi-sited ethnography: Problems and possibilities in the translocation of research methods. New York: Routledge; 2011. p. 16–32.

Google Scholar  

Bentley ME, Pelto GH, Straus WL, Schumann DA, Adegbola C, de la Pena E, et al. Rapid ethnographic assessment: applications in a diarrhea management program. Soc Sci Med. 1988;27(1):107–16.

Article   CAS   PubMed   Google Scholar  

Dixon-Woods M, Martin GP. Does quality improvement improve quality? Future Hosp J. 2016;3(3):191–4.

Leslie M, Paradis E, Gropper MA, Reeves S, Kitto S. Applying ethnography to the study of context in healthcare quality and safety. BMJ Qual Saf. 2014;23(2):99–105.

Article   PubMed   Google Scholar  

Marcus GE. Where have all the tales of fieldwork gone? Ethnos. 2006;71(1):113–22.

Savage J. Ethnography and health care BMJ. 2000;321:1400. https://doi.org/10.1136/bmj.321.7273.1400 .

Waring J, Marshall F, Bishop S. Understanding the occupational and organizational boundaries to safe hospital discharge. Journal of Health Services Research & Policy. 2015;20(1_suppl):35-44. https://doi.org/10.1177/1355819614552512 .

Baim-Lance A, Vindrola-Padros C. Reconceptualising'Impact'through Anthropology's Ethnographic Practices. Anthropol Action. 2015;22(2):5–13.

Latour B. Why Has Critique Run out of Steam? From Matters of Fact to Matters of Concern. Crit Inq. 2004;30(2):225–48.

Zuiderent-Jerak T, Strating M, Nieboer A, Bal R. Sociological refigurations of patient safety; ontologies of improvement and 'acting with' quality collaboratives in healthcare. Soc Sci Med. 2009;69(12):1713–21.

Kitto SC, Sargeant J, Reeves S, Silver I. Towards a sociology of knowledge translation: the importance of being dis-interested in knowledge translation. Adv Health Sci Educ. 2012;17(2):289–99.

Waring J, Allen D, Braithwaite J, Sandall J. Healthcare quality and safety: a review of policy, practice and research. Sociol Health Illn. 2016;38(2):198–215.

Jones L, Pomeroy L, Robert G, Burnett S, Anderson JE, Fulop NJ. How do hospital boards govern for quality improvement? A mixed methods study of 15 organisations in England. BMJ Qual Saf [Internet]. 2017; Available from: http://qualitysafety.bmj.com/content/early/2017/07/07/bmjqs-2016-006433.abstract .

Lucas B, Cooper A, Willson A. The undervalued role of communication in healthcare improvement and its critical contribution to engaging staff and saving lives. J Communication Healthcare. 2021;14(1):5–7.

Lucas B. Getting the improvement habit. BMJ Qual Saf. 2016;25(6):400–3.

Levac D, Colquhoun H, O'Brien KK. Scoping studies: advancing the methodology. Implement Sci. 2010;5(1):1-9.

Arksey H, O'Malley L. Scoping studies: towards a methodological framework. Int J Soc Res Methodol. 2005;8(1):19-32.

Collet JP, Skippen PW, Mosavianpour MK, Pitfield A, Chakraborty B, Hunte G, et al. Engaging pediatric intensive care unit (PICU) clinical staff to lead practice improvement: the PICU participatory action research project (PICU-PAR). Implement Sci. 2014;9:6.

Asan O, Chiou E, Montague E. Quantitative ethnographic study of physician workflow and interactions with electronic health record systems. Int J Ind Ergon. 2015;49:124–30.

Riley R, Coghill N, Montgomery A, Feder G, Horwood J. The provision of NHS health checks in a community setting: an ethnographic account. BMC Health Serv Res. 2015;15:546.

Hjelm M, Holst G, Willman A, Bohman D, Kristensson J. The work of case managers as experienced by older persons (75+) with multi-morbidity - a focused ethnography. BMC Geriatr. 2015;15:168.

Article   PubMed   PubMed Central   CAS   Google Scholar  

Tomnay JE, Bourke L, Fairley CK. Exploring the acceptability of online sexually transmissible infection testing for rural young people in Victoria. Aust J Rural Health. 2014;22(1):40–4.

Van Keer R-L, Deschepper R, Francke AL, Huyghens L, Bilsen J. Conflicts between healthcare professionals and families of a multi-ethnic patient population during critical care: an ethnographic study. Crit Care. 2015;19(1):1–13.

Mylopoulos M, Farhat W. "I can do better": exploring purposeful improvement in daily clinical work. Adv Health Sci Educ. 2015;20(2):371–83.

Spitzmueller MC. Shifting practices of recovery under community mental health reform: A street-level organizational ethnography. Qual Soc Work Res Pract. 2014;13(1):26–48.

Sagasser MH, Fluit CRMG, van Weel C, van der Vleuten CPM, Kramer AWM. How Entrustment Is Informed by Holistic Judgments Across Time in a Family Medicine Residency Program: An Ethnographic Nonparticipant Observational Study. Acad Med. 2017;92(6):792–9.

Edberg M, Cleary S, Simmons LB, Cubilla-Batista I, Andrade EL, Gudger G. Defining the "community": Applying ethnographic methods for a Latino immigrant health intervention. Hum Organ. 2015;74(1):27–41.

Collier A, Sorensen R, Iedema R. Patients' and families' perspectives of patient safety at the end of life: a video-reflexive ethnography study. Int J Qual Health Care. 2016;28(1):66–73.

Mutchler MG, McKay T, McDavitt B, Gordon KK. Using peer ethnography to address health disparities among young urban Black and Latino men who have sex with men. Am J Public Health. 2013;103(5):849–52.

Nelson MM. NICU Culture of Care for Infants with Neonatal Abstinence Syndrome: A Focused Ethnography. Neonatal Network. 2016;35(5):287–96.

Batch M, Windsor C. Nursing casualization and communication: a critical ethnography. J Adv Nurs. 2015;71(4):870–80.

Hughes N. Homelessness, health, and literacy: An institutional ethnographic study of the social organization of health care in Ontario, Canada. In: Dissertation Abstracts International: Section B: The Sciences and Engineering, vol. 78; 2018. (9-B(E)):No-Specified.

Hoare KJ, Buetow S, Mills J, Francis K. Using an emic and etic ethnographic technique in a grounded theory study of information use by practice nurses in New Zealand. J Res Nurs. 2013;18(8):720–31.

Charmaz K, Smith J. Grounded theory. In: Qualitative psychology: A practical guide to research methods, vol. 2; 2003. p. 81–110.

Armstrong N, Brewster L, Tarrant C, Dixon R, Willars J, Power M, et al. Taking the heat or taking the temperature? A qualitative study of a large-scale exercise in seeking to measure for improvement, not blame. Soc Sci Med. 2018;198:157–64.

Boissevain J. Towards a sociology of social anthropology. Theory Soc. 1974;1(2):211–30.

McCabe JL, Holmes D. Reflexivity, critical qualitative research and emancipation: A Foucauldian perspective. J Adv Nurs. 2009;65(7):1518–26.

Nightingale R, Sinha MD, Swallow V. Using focused ethnography in paediatric settings to explore professionals' and parents' attitudes towards expertise in managing chronic kidney disease stage 3-5. BMC Health Serv Res. 2014;14:403.

Stevens J, Schmied V, Burns E, Dahlen HG. Video ethnography during and after caesarean sections: Methodological challenges. J Clin Nurs. 2017;26(13-14):2083–92.

Abdulrehman MS. Reflections on Native Ethnography by a Nurse Researcher. J Transcult Nurs. 2017;28(2):152–8.

Bergenholtz H, Jarlbaek L, Holge-Hazelton B. The culture of general palliative nursing care in medical departments: an ethnographic study. Int J Palliat Nurs. 2015;21(4):193–201.

Belanger E, Rodriguez C, Groleau D, Legare F, MacDonald ME, Marchand R. Patient participation in palliative care decisions: An ethnographic discourse analysis. Int J Qual Stud Health Well Being. 2016;11:32438.

Ranasinghe P. The humdrum of legality and the ordering of an ethic of care. Law Soc Rev. 2014;48(4):709–39.

Johnston MS, Hodge E. ‘Dirt, death and danger? I don't recall any adverse reaction ...': Masculinity and the taint management of hospital private security work. Gend Work Organ. 2014;21(6):546–58.

Babaei S, Taleghani F, Kayvanara M. Compassionate behaviours of clinical nurses in Iran: An ethnographic study. Int Nurs Rev. 2016;63(3):388–94.

Mishra A. 'Trust and teamwork matter': community health workers' experiences in integrated service delivery in India. Global Public Health. 2014;9(8):960–74.

Cribb A. Improvement Science Meets Improvement Scholarship: Reframing Research for Better Healthcare. Health Care Anal. 2018;26(2):109–23.

Boonen MJMH, Vosman FJH, Niemeijer AR, Tinker, tailor, deliberate. An ethnographic inquiry into the institutionalized practice of bar-coded medication administration technology by nurses. Appl Nurs Res. 2017;33:30–5.

Thomas GM, Latimer J. In/exclusion in the clinic: Down's syndrome, dysmorphology and the ethics of everyday medical work. Sociology. 2015;49(5):937–54.

Mackintosh N, Sandall J. The social practice of rescue: the safety implications of acute illness trajectories and patient categorisation in medical and maternity settings. Sociol Health Illness. 2016;38(2):252–69.

Alderson SL, Russell AM, McLintock K, Potrata B, House A, Foy R. Incentivised case finding for depression in patients with chronic heart disease and diabetes in primary care: an ethnographic study. BMJ Open. 2014;4(8):e005146.

Bjornsdottir K. The place of standardisation in home care practice: An ethnographic study. J Clin Nurs. 2014;23(9-10):1411–20.

May M. Turning the board blue: America's epiduralized system of birth. a medical ethnography. In: Dissertation Abstracts International Section A: Humanities and Social Sciences, vol. 76; 2015. (5-A(E)):No-Specified.

Allen D. Lost in translation? 'Evidence' and the articulation of institutional logics in integrated care pathways: from positive to negative boundary object? Sociol Health Illn. 2014;36(6):807–22.

Nilsson L, Eriksen S, Borg C. The influence of social challenges when implementing information systems in a Swedish health-care organisation. J Nurs Manag. 2016;24(6):789–97.

Makaroff KS, Storch J, Pauly B, Newton L. Searching for ethical leadership in nursing. Nurs Ethics. 2014;21(6):642–58.

Pavlish C, Brown-Saltzman K, Jakel P, Fine A. The nature of ethical conflicts and the meaning of moral community in oncology practice. Oncol Nurs Forum. 2014;41(2):130–40.

Gesbeck MM. Negotiating diabetes: Professional diabetes care work in the U.S. In: Dissertation Abstracts International Section A: Humanities and Social Sciences, vol. 77; 2016. (1-A(E)):No-Specified.

Ross C, Rogers C, Duff D. Critical ethnography: An under-used research methodology in neuroscience nursing. Can J Neurosci Nurs. 2016;38(1):4–7.

PubMed   Google Scholar  

Liu W, Manias E, Gerdtz M. Medication communication through documentation in medical wards: knowledge and power relations. Nurs Inq. 2014;21(3):246–58.

Matinga MN, Annegarn HJ, Clancy JS. Healthcare provider views on the health effects of biomass fuel collection and use in rural Eastern Cape, South Africa: an ethnographic study. Soc Sci Med. 1982;2013(97):192–200.

Manias E. Communication relating to family members' involvement and understandings about patients' medication management in hospital. Health Expect. 2015;18(5):850–66.

Goldman J, Reeves S, Wu R, Silver I, MacMillan K, Kitto S. A sociological exploration of the tensions related to interprofessional collaboration in acute-care discharge planning. J Interprof Care. 2016;30(2):217–25.

Fleming DJ. Beyond clinical: The exploration and integration of human connection skills in five residency programs at the university of Arizona. In: Dissertation Abstracts International Section A: Humanities and Social Sciences, vol. 76; 2015. (3-A(E)):No-Specified.

Gealogo GA. "A light in the dark": Development of a conceptual model for person-engaged dementia care. In: Dissertation Abstracts International: Section B: The Sciences and Engineering, vol. 77; 2017. (8-B(E)):No-Specified.

Gillespie BM, Gwinner K, Chaboyer W, Fairweather N. Team communications in surgery-Creating a culture of safety. J Interprof Care. 2013;27(5):387–93.

DeKeyser GF, Engelberg R, Torres N, Curtis JR. Development of a Model of Interprofessional Shared Clinical Decision Making in the ICU: A Mixed-Methods Study. Crit Care Med. 2016;44(4):680–9.

Goldman J, Reeves S, Wu R, Silver I, MacMillan K, Kitto S. Medical residents and interprofessional interactions in discharge: An ethnographic exploration of factors that affect negotiation. J Gen Intern Med. 2015;30(10):1454–60.

Kent F, Francis-Cracknell A, McDonald R, Newton JM, Keating JL, Dodic M. How do interprofessional student teams interact in a primary care clinic? A qualitative analysis using activity theory. Adv Health Sci Educ. 2016;21(4):749–60.

Milne J, Greenfield D, Braithwaite J. An ethnographic investigation of junior doctors' capacities to practice interprofessionally in three teaching hospitals. J Interprof Care. 2015;29(4):347–53.

Reeves S, McMillan SE, Kachan N, Paradis E, Leslie M, Kitto S. Interprofessional collaboration and family member involvement in intensive care units: emerging themes from a multi-sited ethnography. J Interprof Care. 2015;29(3):230–7.

Collier A, Wyer M. Researching Reflexively With Patients and Families: Two Studies Using Video-Reflexive Ethnography to Collaborate With Patients and Families in Patient Safety Research. Qual Health Res. 2016;26(7):979–93.

Nastasi BK, Schensul JJ, Schensul SL, Mekki-Berrada A, Pelto PJ, Maitra S, et al. A model for translating ethnography and theory into culturally constructed clinical practices. Cult Med Psychiatry. 2015;39(1):92–120.

Nugus P, Forero R, McCarthy S, McDonnell G, Travaglia J, Hilman K, et al. The emergency department "carousel": an ethnographically-derived model of the dynamics of patient flow. Int Emerg Nurs. 2014;22(1):3–9.

MacKichan F, Brangan E, Wye L, Checkland K, Lasserson D, Huntley A, et al. Why do patients seek primary medical care in emergency departments? An ethnographic exploration of access to general practice. BMJ Open. 2017;7(4):e013816.

Nicolini D. Zooming in and out: Studying practices by switching theoretical lenses and trailing connections. Organ Stud. 2009;30(12):1391–418.

Tietbohl CK, Rendle KAS, Halley MC, May SG, Lin GA, Frosch DL. Implementation of Patient Decision Support Interventions in Primary Care: The Role of Relational Coordination. Med Decis Mak. 2015;35(8):987–98.

Ash JS, Chase D, Wiesen JF, Murphy EV, Marovich S. Studying Readiness for Clinical Decision Support for Worker Health Using the Rapid Assessment Process and Mixed Methods Interviews. AMIA Annu Symp proc AMIA Symposium. 2016;2016:285–94.

Balka E, Tolar M, Coates S, Whitehouse S. Socio-technical issues and challenges in implementing safe patient handovers: insights from ethnographic case studies. Int J Med Inform. 2013;82(12):e345–57.

Borycki EM, Kushniruk AW. Use of Techno-Anthropologic Approaches in Studying Technology--induced Errors. Stud Health Technol Inform. 2015;215:129–41.

Dixon-Woods M, Redwood S, Leslie M, Minion J, Martin GP, Coleman JJ. Improving quality and safety of care using “technovigilance”: an ethnographic case study of secondary use of data from an electronic prescribing and decision support system. Milbank Q. 2013;91(3):424–54.

Wright A, Sittig DF, Ash JS, Erickson JL, Hickman TT, Paterno M, et al. Lessons learned from implementing service-oriented clinical decision support at four sites: A qualitative study. Int J Med Inform. 2015;84(11):901–11.

Pfadenhauer M, Dukat C. Robot caregiver or robot-supported caregiving? The performative deployment of the social robot PARO in dementia care. Int J Soc Robot. 2015;7(3):393–406.

Jensen S. Clinical Simulation: For what and how can it be used in design and evaluation of health IT. Stud Health Technol Inform. 2015;215:217–28.

McNall MA, Welch VE, Ruh KL, Mildner CA, Soto T. The use of rapid-feedback evaluation methods to improve the retention rates of an HIV/AIDS healthcare intervention. Eval Program Plann. 2004;27(3):287–94.

Grant S, Mesman J, Guthrie B. Spatio-temporal elements of articulation work in the achievement of repeat prescribing safety in UK general practice. Soc Health Illness. 2016;38(2):306–24.

Grant S, Guthrie B. Efficiency and thoroughness trade-offs in high-volume organisational routines: an ethnographic study of prescribing safety in primary care. BMJ Qual Saf. 2018;27(3):199–206.

Mendenhall E, Yarris K, Kohrt BA. Utilization of standardized mental health assessments in anthropological research: Possibilities and pitfalls. Cult Med Psychiatry. 2016;40(4):726–45.

Street A, Coleman S. Introduction: real and imagined spaces. Space Cult. 2012;15(1):4–17.

Sandelowski M. A matter of taste: evaluating the quality of qualitative research. Nurs Inq. 2015;22(2):86–94.

Vindrola-Padros C, Vindrola-Padros B. Quick and dirty? A systematic review of the use of rapid ethnographies in healthcare organisation and delivery. BMJ Qual Saf. 2018;27(4):321–30.

Download references

Acknowledgements

The authors wish to thank Lorelei Jones, Natalie Armstrong, Justin Waring and Bill Lucas for their insightful comments and direction in the undertaking of this work.

This paper is independent research funded by the National Institute for Health Research CLAHRC North Thames. The views expressed in this publication are those of the author(s) and not necessarily those of the National Institute for Health Research or the Department of Health and Social Care.

NJF is an NIHR Senior Investigator. GB is supported by the Health Foundation’s grant to the University of Cambridge for The Healthcare Improvement Studies Institute.

Author information

Authors and affiliations.

Department of Applied Health Research, UCL, London, UK

Georgia B. Black, Sandra van Os, Samantha Machen & Naomi J. Fulop

You can also search for this author in PubMed   Google Scholar

Contributions

NJF and GB led the development and conceptualization of this scoping review and provided guidance on methods and design of the scoping review. GB, SVO and SM made contributions to study search, study screening, and all data extraction work. All authors analysed the data. All authors contributed to the writing and editing of the paper, and all authors have read and approved the manuscript.

Corresponding author

Correspondence to Georgia B. Black .

Ethics declarations

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

The authors have no competing interests to declare.

Additional information

Publisher’s note.

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

The original online version of this article was revised: due to incorrect figure 1 and the number of included papers need to be changed from "283" to "274".

Supplementary Information

Additional file 1., additional file 2., additional file 3., additional file 4., 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.

Black, G.B., van Os, S., Machen, S. et al. Ethnographic research as an evolving method for supporting healthcare improvement skills: a scoping review. BMC Med Res Methodol 21 , 274 (2021). https://doi.org/10.1186/s12874-021-01466-9

Download citation

Received : 21 July 2021

Accepted : 14 October 2021

Published : 05 December 2021

DOI : https://doi.org/10.1186/s12874-021-01466-9

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

  • Ethnography
  • Qualitative research
  • Healthcare improvement

BMC Medical Research Methodology

ISSN: 1471-2288

ethnographic case study methodology

ethnographic case study methodology

How to... Use ethnographic methods & participant observation

Find out how to use ethnographic research methods and participant observation in our detailed guide.

On this page

What are ethnographic methods, organisational ethnography, what is participant observation , analysing, theorising and writing up.

Ethnographic methods are a research approach where you look at people in their cultural setting, with the goal of producing a narrative account of that particular culture, against a theoretical backdrop. As part of this you will look at:

  • Deeds done as well as words used
  • How they interact with one another, and with their social and cultural environment
  • What is  not  said as much as what is said
  • Language, and symbols, rituals and shared meanings that populate their world

Ethnography is a study of culture, therefore, organisational ethnography looks at the culture of organisations.

Organisational culture exists within the minds of the people who make up that organisation, while organisational ethnography is concerned with settings within which social relations take place between actors who are set on particular goals.

This culture evolves over time, contains dominant cultures and subcultures, and is subject to its own rules, rites, myths and symbols.

History of ethnographic methods

Ethnography has its origins in social anthropology, and in particular, the work of Malinowski whose seminal text  Argonauts of the Western Pacific  describes his experience of living for a long time with South Pacific islanders, and counsels the anthropologist to spend at least a year in the field, to learn the language, and to live as one of the population which he or she studies.

It was taken over by sociology in the 1930s when the Chicago school studied "deviant subcultures" in urban America in the great depression.

Early ethnographers were criticized for their detached stance, particularly by feminist anthropologists, but recent adaptations of the method use it in action research, where the study population itself becomes involved in the request for information and meaning.

Research parameters

Ethnographic methods are qualitative, inductive, exploratory and longitudinal. They achieve a thick, rich description over a relatively small area.

As the researcher, it is best if you conduct your data gathering on an iterative basis, with you taking on a "reflexive" role – in other words observing, reflecting, building up a theory and then going back into the field and testing it.

This process of testing is essential, because of the inevitable element of subjectivity in a research method where you, the researcher, is the instrument.

There are a number of practical considerations with ethnographic methods, such as:

  • Time.  Studies are time-consuming to complete. If you are looking at making ethnography one of your approaches for a dissertation, will you have sufficient time before the completion date? If part of a major research project, will the project bear the costs?
  • Place.  You need to make sure that you can get the cooperation of the organisation you wish to observe and decide whether you want to look at the whole organisation, one part of it or a cross-section.

Data collection methods and triangulation

Most ethnographic research makes considerable use of participant observation, usually triangulated with interviews and/or ordinary "informal" conversations.

Triangulation is particularly important as one method on its own is not usually reliable.

You can also gain a lot of information from other sources, such as:

  • Written documents , e.g. e-mails, policy documents, meeting minutes, organisation charts, reports, procedural manuals, "official" corporate material such as an intranet, brochures, press releases, advertising, web pages, annual report.
  • Corporate events  like the annual staff conference and Christmas party, etc.
  • Branding  – logo and how it is applied, slogan, etc. Branding is a particularly strong use of symbolism.
  • Site location , built environment, design, etc.

Another method used is that of the diary, which participants are required to complete (you will also be completing a diary as part of your participant observation.

This may either have set categories as in structured observation, or the participant may be required to keep a record of their experiences (for example, their reactions to a training course) or of what they do.

Participant observation is one of the main ethnographic data collection methods.

The essence of participant observation is that you, as the researcher, observe the subject of research, either by participating directly in the action, as a member of the study population, or as a "pure" observer, in which case you do not participate in the action but are still present on the scene, for example observing workers in a manufacturing plant or discussants in the board room.

In either case, you observe, note, record, describe, analyse, and interpret people and their interactions, and related events, with the objective of obtaining a systematic account of behaviour and idea systems of a given community, organisation or institution.

Why use participant observation?

Like other ethnographic methods, participant observation is very much based on the classic methods used in early anthropology, by Malinowski and others as they studied particular populations, often for years at a time, taking detailed notes.

Participant observation is usually inductive, and carried out as part of an exploratory research phase, with the view of forming hypotheses from the data. It is often connected with the  grounded theory  method, according to which researchers revisit the research territory with deeper and deeper knowledge.

The strength of participant observation is its ability to describe depth (thick description) and to help understand human behaviour.

Researcher roles

There is a continuum in observation techniques between the covert and the overt observer, and the observer who participates completely in the activity and the one who is purely a "fly on the wall".

There are problems with all these approaches, but the ideal is to ensure that the maximum amount of information is gained whilst at the same time retaining the maximum distance in order to ensure researcher objectivity.

Which role is adopted would depend on the subject being researched, for example:

  • Complete participant.  There are obvious ethical considerations of being part of a group and not revealing your role as a researcher who will subsequently write up the research study, but in some circumstances revealing your role might prejudice the research, particularly if the subject concerns something delicate such as the consumption of alcohol or drugs.
  • Complete observer.  This might be appropriate to a situation where the subject is relatively large-scale, for example observing people in a shopping mall or in a supermarket, or where the revelation of the role might destroy the dynamics of small group behaviour, as for example watching the behaviour of groups of shoppers.
  • Observer as participant.  The disadvantage here is that although you participate in some way in the activity, you lose the emotional involvement, but the advantage is that you can concentrate on your role as a researcher. It might be used if you were, for example, observing people on a training course, or users of electronic courseware, where it was very important to understand the reactions and mental processes of the participants rather than what they do.
  • Participant as observer.  The advantage of participating is that you become fully part of the group, and you can experience directly what your subjects are experiencing. It is particularly useful when for example you need to understand work practices or job roles.

There are other possible roles for the observer:

  • As  facilitator  and  change agent , when you become involved not merely as a participant but as someone who seeks to help subjects change some aspect of their world, for example in action research.
  • As a  narrator , describing what you have witnessed from a position of authority.

Structured observation

Structured observation differs from participant observation in that it is more detached, more systematic, and what is observed often has a more mechanical quality. It is also a quantitative as opposed to a qualitative technique, concerned with quantifying behaviour as opposed to obtaining a rich description.

Advantages and disadvantages

Participant observation is not without its detractors and is seen to have a number of advantages and disadvantages:

  • It can provide very rich data and can be particularly good at revealing facets of human behaviour.
  • It does not rely on the words of the actors themselves, so is not dependent on people's ability to verbalize, and provides a source other than their own testimony.
  • Issues of time and researcher objectivity can be met by careful use of sampling, whereby observation is confined to particular locations and times.
  • All research techniques have inbuilt problems of bias: for example, the interviewer effect, and the difficulty of formulating careful survey questions.

Disadvantages

  • Because the researcher is the instrument, there may be difficulty in maintaining the necessary objectivity.
  • Good participant observation takes up a lot of time.
  • It requires intimacy and an invasion of privacy which may be disruptive both to the research process and also to the organisation itself. On the other hand, concealing one's identity is ethically questionable as it involves a deception.
  • Observer bias: the observer's own views and personal beliefs may impinge upon observations.

However, the best way of using participant observation in a useful and responsible way is to triangulate it with other approaches.

Applications to management research

Participant observation is based on the social sciences, particularly social anthropology and on the premise that you go and study a different, and often remote culture.

The appeal to management research is that it can study the culture of an organisation in depth. However, in many cases it is simply not practical to immerse oneself for months at a time: the cost would be too great, the organisation may not be willing, and one cannot actually live with the workers. For this reason,  time sampling  is often adopted, where the times at which observation takes place are carefully selected.

Use in market research

Participant observation is particularly useful in market research. It is a natural technique as both are concerned with human behaviour. It can be a good method when:

  • The subject under study is easily observable and occurs in public
  • It concerns a social process or mass activity, such as the disposal of household waste
  • The processes are subconscious, for example in a study of in-store music
  • It would not be desirable or easy for consumers to interact with the researcher, for example with very young children.

Data collection

Observations should be recorded as far as is possible on the day of the fieldwork, in diary form, and should comprise the following:

  • Time of day
  • Actors present
  • Sequence of events, and any interruptions.
  • Secondary observations  in the form of any statements by others about what you observed.
  • Experiential data  as relating to your own state of mind, emotions and any reflections.
  • Circumstantial and background data  about the organisation, key roles etc.

Analysis of unstructured data

What distinguishes the analysis of ethnographically generated data is that the research process is inductive and iterative.

Unlike the neatly linear trajectory of some other research, when you construct an instrument to prove a theory and do not analyse until you have collected all the data, in ethnographic research data collection and analysis may be simultaneous, while theories are formed on the basis of some data and then tested and refined against further data. This process is known as  analytic induction .

When you begin to collect data, you will find very soon that you get a lot. This is the time to begin an initial analysis. As you start the coding process, begin to look for groupings, based on frequency and patterns of and in the data. As you refine your  coding structure , check your assumptions carefully. Eventually, you will reach a point where you are relatively confident of your coding structure and you can begin to use it as a way of organising your data.

There are a number of  software packages  –  NVivo, QSR NUD.IST  and  The Ethnograph  for example – that can help here, or you may prefer to use an ordinary office package such as Word or Excel. Some of the software packages also offer modelling facilities.

Whatever method you use, at this stage patterns will begin to emerge from which you will be able to build theory.

Analysis of structured data

The analysis of structured observation data is different in that the coding schedule is established before the start of data collection. In this case you either:

  • establish your own headings, which should be consistent with your research questions
  • follow an existing "off the shelf" coding schedule
  • use a combination of these approaches, modifying an existing schedule and perhaps putting in some of your own headings.

The fact that data are situation specific and not easy to replicate, together with the possibility of observer bias, are threats to validity with unstructured observation.

These threats can be dealt with by:

  • Checking the observations, and interpretations of them, with participants, as a form of triangulation.
  • Checking the coding structure, which can be done by the researcher checking against emerging theory, and other researchers coding the data to see if they come up with similar coding structures.
  • "Perspicacity" – the ability to abstract from the data general principles that can throw light on other similar situations.

Theory building

The literature review is commonly done at the beginning of the research process. But with ethnographic research, it often follows (at least some) data collection and analysis – because it is connected with theory building.

In ethnographic research, the researcher is often compared with a journalist researching a story and looking for promising lines of enquiry. As the data are being collected and patterns start to emerge, so may interesting lines of enquiry on which theories can be built.

The objective of the theory is not to predict, but to explain, to look for contextual structures and to provide a context for events, conversations and descriptions. You are providing an explanatory framework for the phenomena which you have been observing.

As indicated above, once you have formulated a theory you need to check it against the data, and check the data against itself – how valid is it?

The theory also needs to be situated in the relevant literature, and have its own theoretical context.

For a dissertation , you should follow the guidelines of your own university and check out other dissertations which have used similar research techniques.

A traditional approach, however, is introduction, literature review, philosophical approach and methodology, findings, analysis, discussion and conclusion.

For a journal article , you are best advised to look carefully at other examples of articles written for scholarly journals, particularly ones in which you are thinking of publishing.

Ethnographic research allows us to regard and represent the actors as creators as well as executants of their own meanings. The very way in which they tell us about what they do tells the researcher a great deal about what is meaningful for and in the research. It adds richness and texture to the experience of conducting research.

Related topics

Using mixed methods research.

In our experience, many editors are particularly pleased to receive submissions that combine qualitative and quantitative research. Find out more about this "mixed methods" approach.

Make your research easy to find with SEO

Help your article gain attention with clever use of search engine optimisation (SEO) at the writing stage.

Proofreading

In this guide, we explain what you should look for at the proofing stage.

  • Privacy Policy

Research Method

Home » Ethnographic Research -Types, Methods and Guide

Ethnographic Research -Types, Methods and Guide

Table of Contents

Ethnographic Research

Ethnographic Research

Definition:

Ethnographic research is a qualitative research method used to study and document the culture, behaviors, beliefs, and social interactions of a particular group of people. It involves direct observation and participation in the daily life and activities of the group being studied, often for an extended period of time.

Ethnographic Study

An ethnographic study is a research method that involves the detailed and systematic study of a particular group, culture, or community. Ethnographic studies seek to understand the beliefs, values, behaviors, and social dynamics of a group through direct observation and participation in their daily life.

Ethnographic Research vs Ethnographic Study

here’s a table comparing ethnographic study and ethnographic research:

AspectEthnographic StudyEthnographic Research
Qualitative research methodQualitative research method
Study of a particular group, culture, or communityResearch on a particular group, culture, or community
To understand the culture, beliefs, behaviors, and social interactions of the group being studiedTo document and analyze the culture, beliefs, behaviors, and social interactions of the group being studied
Participant observation, interviews, surveys, and document analysisParticipant observation, interviews, surveys, and document analysis
In-depth and systematic study of the group over an extended period of timeCollection of data through various techniques and analysis of the data collected
Develop a holistic and nuanced understanding of the community being studiedDocument and provide insights into the culture, practices, and social dynamics of the community being studied
Used to inform policy decisions or address social issues related to specific communitiesUsed to explore and document the diversity of human cultures and societies or to inform policy decisions or address social issues related to specific communities

While there are some differences between the two, they are similar in that they both use qualitative research methods to study a particular group, culture, or community. The main difference is that an ethnographic study involves the researcher spending an extended period of time within the community being studied in order to develop a deep understanding, while ethnographic research is focused on documenting and analyzing the culture, beliefs, behaviors, and social interactions of the group being studied.

Ethnographic Research Types

Ethnographic research can be divided into several types based on the focus of the study and the research objectives. Here are some common types of ethnographic research:

Classic Ethnography

This type of ethnographic research involves an extended period of observation and interaction with a particular community or group. The researcher aims to understand the community’s culture, beliefs, practices, and social structure by immersing themselves in the community’s daily life.

Autoethnography

Autoethnography involves the researcher using their own personal experiences to gain insights into a particular community or culture. The researcher may use personal narratives, diaries, or other forms of self-reflection to explore the ways in which their own experiences relate to the culture being studied.

Participatory Action Research

Participatory action research involves the researcher working collaboratively with members of a particular community or group to identify and address social issues affecting the community. The researcher aims to empower community members to take an active role in the research process and to use the findings to effect positive change.

Virtual Ethnography

Virtual ethnography involves the use of online or digital media to study a particular community or culture. The researcher may use social media, online forums, or other digital platforms to observe and interact with the group being studied.

Critical Ethnography

Critical ethnography aims to expose power imbalances and social inequalities within a particular community or culture. The researcher may use their observations to critique dominant cultural narratives or to identify opportunities for social change.

Ethnographic Research Methods

Some common ethnographic research methods include:

Participant Observation

This involves the researcher directly observing and participating in the daily life and activities of the group being studied. This technique helps the researcher gain an in-depth understanding of the group’s behavior, culture, and social dynamics.

Ethnographic researchers use interviews to gather information about the group’s beliefs, values, and practices. Interviews may be formal or informal and can be conducted one-on-one or in group settings.

Surveys can be used to collect data on specific topics, such as attitudes towards a particular issue or behavior patterns. Ethnographic researchers may use surveys as a way to gather quantitative data in addition to qualitative data.

Document Analysis

This involves analyzing written or visual documents produced by the group being studied, such as newspapers, photographs, or social media posts. Document analysis can provide insight into the group’s values, beliefs, and practices.

Field Notes

Ethnographic researchers keep detailed field notes of their observations and interactions with the group being studied. These notes help the researcher organize their thoughts and observations and can be used to analyze the data collected.

Focus Groups

Focus groups are group interviews that allow the researcher to gather information from multiple people at once. This technique can be useful for exploring shared beliefs or experiences within the group being studied.

Ethnographic Research Data Analysis Methods

Ethnographic research data analysis methods involve analyzing qualitative data collected from observations, interviews, and other sources in order to identify patterns, themes, and insights related to the research question.

Here are some common data analysis methods used in ethnographic research:

Content Analysis

This involves systematically coding and categorizing the data collected from field notes, interviews, and other sources. The researcher identifies recurring themes, patterns, and categories in the data and assigns codes or labels to each one.

Narrative Analysis

This involves analyzing the stories and narratives collected from participants in order to understand how they construct and make sense of their experiences. The researcher looks for common themes, plot structures, and rhetorical strategies used by participants.

Discourse Analysis

This involves analyzing the language and communication practices of the group being studied in order to understand how they construct and reproduce social norms and cultural meanings. The researcher looks for patterns in the use of language, including metaphors, idioms, and other linguistic devices.

Comparative Analysis

This involves comparing data collected from different groups or communities in order to identify similarities and differences in their cultures, behaviors, and social structures. The researcher may use this analysis to generate hypotheses about why these differences exist and what factors may be contributing to them.

Grounded Theory

This involves developing a theoretical framework based on the data collected during the research process. The researcher identifies patterns and themes in the data and uses these to develop a theory that explains the social phenomena being studied.

How to Conduct Ethnographic Research

To conduct ethnographic research, follow these general steps:

  • Choose a Research Question: Identify a research question that you want to explore. It should be focused and specific, but also open-ended to allow for flexibility and exploration.
  • Select a research site: Choose a site or group that is relevant to your research question. This could be a workplace, a community, a social movement, or any other social setting where you can observe and interact with people.
  • Obtain ethical clearance: Obtain ethical clearance from your institution or organization before beginning your research. This involves ensuring that your research is conducted in an ethical and responsible manner, and that the privacy and confidentiality of participants are protected.
  • Conduct observations: Observe the people in your research site and take detailed notes. This involves being present and engaged in the social setting, participating in activities, and taking note of the behaviors, interactions, and social norms that you observe.
  • Conduct interviews : Conduct interviews with people in the research site to gain deeper insights into their experiences, perspectives, and beliefs. This could involve structured or semi-structured interviews, focus groups, or other forms of data collection.
  • Analyze data: Analyze the data that you have collected, looking for themes and patterns that emerge. This involves immersing yourself in the data and interpreting it within the social and cultural context of the research site.
  • Write up findings: Write up your findings in a clear and concise manner, using quotes and examples to illustrate your key points. This may involve creating narratives, tables, or other visual representations of your findings.
  • Reflect on your process: Reflect on your process and methods, thinking about what worked well and what could be improved for future research.

When to Use Ethnographic Research

Here are some situations where ethnographic research may be particularly appropriate:

  • When exploring a new topic: Ethnographic research can be useful when exploring a topic that has not been well-studied before. By engaging with members of a particular group or community, researchers can gain insights into their experiences and perspectives that may not be visible from other research methods.
  • When studying cultural practices: Ethnographic research is particularly useful when studying cultural practices and beliefs. By immersing themselves in the cultural context being studied, researchers can gain a deeper understanding of the ways in which cultural practices are enacted, maintained, and transmitted.
  • When studying complex social phenomena: Ethnographic research can be useful when studying complex social phenomena that cannot be easily understood through quantitative methods. By observing social interactions and behaviors, researchers can gain insights into the ways in which social norms and structures are created and maintained.
  • When studying marginalized communities: Ethnographic research can be particularly useful when studying marginalized communities, as it allows researchers to give voice to members of these communities and understand their experiences and perspectives.

Overall, ethnographic research can be a useful research approach when the goal is to gain a deep understanding of a particular group or community and their cultural practices, beliefs, and experiences. It is a flexible and adaptable research method that can be used in a variety of research contexts.

Applications of Ethnographic Research

Ethnographic research has many applications across a wide range of fields and disciplines. Some of the key applications of ethnographic research include:

  • Informing policy and practice: Ethnographic research can provide valuable insights into the experiences and perspectives of marginalized or underrepresented groups, which can inform policy and practice in fields such as health care, education, and social services.
  • Developing theories and concepts: Ethnographic research can contribute to the development of theories and concepts in social and cultural anthropology, sociology, and other disciplines, by providing detailed and nuanced accounts of social and cultural phenomena.
  • Improving product design and marketing: Ethnographic research can be used to understand consumer behavior and preferences, which can inform the design and marketing of products and services.
  • Studying workplace culture: Ethnographic research can provide insights into the norms, values, and practices of organizations, which can inform efforts to improve workplace culture and employee satisfaction.
  • Examining social movements: Ethnographic research can be used to study the practices, beliefs, and experiences of social movements, which can inform efforts to understand and address social and political issues.
  • Studying healthcare practices: Ethnographic research can provide insights into healthcare practices and patient experiences, which can inform efforts to improve healthcare delivery and patient outcomes.

Examples of Ethnographic Research

Here are some real-time examples of ethnographic research:

  • Anthropological study of a remote indigenous tribe: Anthropologists often use ethnographic research to study remote indigenous tribes and gain insights into their culture, beliefs, and practices. For example, an anthropologist may live with a tribe for an extended period of time, observing and participating in their daily activities, and conducting interviews with members of the community.
  • Study of workplace culture: Ethnographic research can be useful in studying workplace culture and understanding the dynamics of the organization. For example, an ethnographer may observe and interview employees in a particular department or team to gain insights into their work practices, communication styles, and social dynamics.
  • Study of consumer behavior: Ethnographic research can be useful in studying consumer behavior and understanding how people interact with products and services. For example, an ethnographer may observe and interview consumers as they use a particular product, such as a new smartphone or fitness tracker, to gain insights into their behaviors and preferences.
  • Study of health care practices: Ethnographic research can be useful in studying health care practices and understanding how patients and providers interact within the health care system. For example, an ethnographer may observe and interview patients and providers in a hospital or clinic to gain insights into their experiences and perspectives.
  • Study of social movements: Ethnographic research can be useful in studying social movements and understanding how they emerge and evolve over time. For example, an ethnographer may observe and interview participants in a protest movement to gain insights into their motivations and strategies.

Purpose of Ethnographic Research

The purpose of ethnographic research is to provide an in-depth understanding of a particular group or community, including their cultural practices, beliefs, and experiences. This research approach is particularly useful when the research question is exploratory and the goal is to generate new insights and understandings. Ethnographic research seeks to understand the experiences, perspectives, and behaviors of the participants in their natural setting, without imposing the researcher’s own biases or preconceptions.

Ethnographic research can be used to study a wide range of topics, including social movements, workplace culture, consumer behavior, and health care practices, among others. The researcher aims to understand the social and cultural context of the group or community being studied, and to generate new insights and understandings that can inform future research, policy, and practice.

Overall, the purpose of ethnographic research is to gain a deep understanding of a particular group or community, with the goal of generating new insights and understandings that can inform future research, policy, and practice. Ethnographic research can be a valuable research approach in many different contexts, particularly when the goal is to gain a rich, contextualized understanding of social and cultural phenomena.

Advantages of Ethnographic Research

Ethnographic research has several advantages that make it a valuable research approach in many different fields. Here are some of the advantages of ethnographic research:

  • Provides in-depth and detailed information: Ethnographic research involves direct observation of the group or community being studied, which allows researchers to gain a detailed and in-depth understanding of their beliefs, practices, and experiences. This type of information cannot be obtained through other research methods.
  • Offers a unique perspective: Ethnographic research allows researchers to see the world from the perspective of the group or community being studied. This can provide unique insights into the ways in which different cultural practices and beliefs are constructed and maintained.
  • Promotes cultural understanding: Ethnographic research can help to promote cultural understanding and reduce stereotypes by providing a more nuanced and accurate picture of different cultures and communities.
  • Allows for flexibility: Ethnographic research is a flexible research approach that can be adapted to fit different research contexts and questions. Researchers can adjust their methods based on the needs of the group being studied and the research goals.
  • Generates rich and diverse data: Ethnographic research generates rich and diverse data through a combination of observation, interviews, and other methods. This allows researchers to analyze different aspects of the group or community being studied and identify patterns and themes in the data.
  • Supports theory development: Ethnographic research can support theory development by providing empirical data that can be used to test and refine theoretical frameworks.

Limitations of Ethnographic Research

Ethnographic research has several limitations that researchers should consider when selecting this research approach. Here are some of the limitations of ethnographic research:

  • Limited generalizability: Ethnographic research typically involves studying a small and specific group or community, which limits the generalizability of the findings to other contexts or populations.
  • Time-consuming: Ethnographic research is a time-consuming process that requires a significant investment of time and resources. Researchers must spend time observing and interacting with the group being studied, which may not be feasible in all research contexts.
  • Subjectivity: Ethnographic research relies on the researcher’s interpretation and analysis of the data collected, which may introduce subjective bias into the research findings.
  • Limited control: Ethnographic research involves studying a group or community in their natural setting, which limits the researcher’s control over the research context and the behavior of the participants.
  • Ethical concerns: Ethnographic research can raise ethical concerns, particularly when studying marginalized or vulnerable populations. Researchers must be careful to ensure that they do not harm or exploit the participants in the research process.
  • Limited quantitative data: Ethnographic research typically generates qualitative data, which may limit the types of analysis that can be conducted and the types of conclusions that can be drawn.

About the author

' src=

Muhammad Hassan

Researcher, Academic Writer, Web developer

You may also like

Basic Research

Basic Research – Types, Methods and Examples

Correlational Research Design

Correlational Research – Methods, Types and...

Transformative Design

Transformative Design – Methods, Types, Guide

Survey Research

Survey Research – Types, Methods, Examples

Focus Groups in Qualitative Research

Focus Groups – Steps, Examples and Guide

Questionnaire

Questionnaire – Definition, Types, and Examples

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

Logo of nihpa

Research MethodologyOverview of Qualitative Research

Qualitative research methods are a robust tool for chaplaincy research questions. Similar to much of chaplaincy clinical care, qualitative research generally works with written texts, often transcriptions of individual interviews or focus group conversations and seeks to understand the meaning of experience in a study sample. This article describes three common methodologies: ethnography, grounded theory, and phenomenology. Issues to consider relating to the study sample, design, and analysis are discussed. Enhancing the validity of the data, as well reliability and ethical issues in qualitative research are described. Qualitative research is an accessible way for chaplains to contribute new knowledge about the sacred dimension of people's lived experience.

INTRODUCTION

Qualitative research is, “the systematic collection, organization, and interpretation of textual material derived from talk or conversation. It is used in the exploration of meanings of social phenomena as experienced by individuals themselves, in their natural context” ( Malterud, 2001 , p. 483). It can be the most accessible means of entry for chaplains into the world of research because, like clinical conversations, it focuses on eliciting people's stories. The stories can actually be expressed in almost any medium: conversations (interviews or focus groups), written texts (journal, prayers, or letters), or visual forms (drawings, photographs). Qualitative research may involve presenting data collected from a single person, as in a case study ( Risk, 2013 ), or from a group of people, as in one of my studies of parents of children with cystic fibrosis (CF) ( Grossoehme et al., 2013 ). Whole books are devoted to qualitative research methodology and, indeed, to the individual methods themselves. This article is intended to present, in rather broad brushstrokes, some of the “methods of choice” and to suggest some issues to consider before embarking on a qualitative research project. Helpful texts are cited to provide resources for more complete information.

Although virtually anything may be data, spoken mediums are the most common forms of collecting data in health research, so the focus of this article will mainly be on interviews and to a lesser extent, focus groups. Interviews explore experiences of individuals, and through a series of questions and answers, the meaning individuals give to their experiences ( Tong, Sainsbury, & Craig, 2007 ). They may be “structured” interviews, in which an interview guide is used with pre-determined questions from which no deviation is permitted by the interviewer, or semi-structured interviews, in which an interview guide is used with pre-determined questions and potential follow-up questions. The latter allows the interviewer to pursue topics that arise during the interview that seem relevant ( Cohen & Crabtree, 2006 ). Writing good questions is harder than it appears! In my first unit of CPE, the supervisor returned verbatims, especially our early efforts, with “DCFQ” written in the margin, for “direct, closed, factual question.” We quickly learned to avoid DCFQs in our clinical conversations because they did not create the space for reflection on illness and the sacred the way open-ended questions did. To some extent, writing good open-ended questions that elicit stories can come more readily to chaplains, due perhaps to our training, than to investigators from other disciplines. This is not to say writing an interview guide is easy or an aspect of research that can be taken lightly, as the quality of the data you collect, and hence the quality of your study, depends on the quality of your interview questions.

Data may also be collected using focus groups. Focus groups are normally built around a specific topic. They almost always follow a semi-structured format and include open discussion of responses among participants, which may range from four to twelve people ( Tong et al., 2007 ). They provide an excellent means to gather data on an entire range of responses to a topic, or on the social interactions between participants, or to clarify a process. Once the data are collected, the analytic approach is typically similar to that of interview data.

Qualitative investigators are not disinterested outsiders who merely observe without interacting with participants, but affect and are affected by their data. The investigator's emotions as they read participants' narratives are data to be included in the study. Simply asking “research” questions can itself be a chaplaincy intervention: what we ask affects the other person and can lead them to reflect and change ( Grossoehme, 2011 ). It is important to articulate our biases and understand how they influence us when we collect and analyze data. Qualitative research is often done by a small group of researchers, especially the data coding. This minimizes the bias of an individual investigator. Inevitably, two or more people will code passages differently at times. It is important to establish at the outset how such discrepancies will be handled.

Ensuring Rigor, Validity, and Reliability

Some people do not think qualitative research is not very robust or significant. This attitude is due, in part, to the poor quality of some early efforts. Increasingly, however, qualitative studies have improved in rigor, and reviewers of qualitative manuscripts expect investigators to have addressed problematic issues from the start of the project. Two important areas are validity and reliability. Validity refers to whether or not the final product (usually referred to a “model”) truly portrays what it claims to portray. If you think of a scale on which you weigh yourself, you want a valid reading so that you know your correct weight. Reliability refers to the extent to which the results are repeatable; if someone else repeated this study, would they obtain the same result? To continue the scale analogy, a reliable scale gives the same weight every time I step on it. A scale can be reliable without being valid. The scale could reliably read 72 pounds every time I step on it, but that value is hardly correct, so the measure is not valid.

Swinton and Mowat (2006) discussed ensuring the “trustworthiness” of the data. N narrative data which are “rich” in their use of metaphor and description, and which express deeper levels of meaning and nuance compared to everyday language are likely to yield a trustworthy final model because the investigators have done a credible job of completely describing and understanding the topic that is under study. Validity is also enhanced by some methodologies, such as grounded theory, which use participants' own words to name categories and themes, instead of using labels given by the investigator. The concept of “member checking” also enhances validity. Once the analyses are complete and a final model has been developed, these findings are shown to all or some of the participants (the members) who are invited to check the findings and give feedback. Do they see themselves in the words or conceptual model that is presented? Do they offer participants a new insight, or do they nod agreement without really reengaging the findings?

Reliability

One means of demonstrating reliability is to document the research decisions made along the way, as they were made, perhaps in a research diary ( Swinton & Mowat, 2006 ). Qualitative methodologies accept that the investigator is part of what is being studied and will influence it, and that this does not devalue a study but, in fact, enhances it. Simply deciding what questions to ask or not ask, and who you ask them to (and not) reflect certain decisions that should be consciously made and documented. Another researcher should be able to understand what was done and why from reading the research diary.

ETHNOGRAPHIC RESEARCH

Elisa Sobo (2009) defines ethnography as the presentation of, “… a given group's conceptual world, seen and experienced from the inside” (p. 297). Ethnography answers the question, “what's it like to be this person?” One example of this kind of study comes from the work of Fore and colleagues ( Fore, Goldenhar, Margolis, & Seid, 2013 ). In order to design tools that would enable clinicians and persons with pediatric inflammatory bowel disease (IDB) to work together more efficiently, an ethnographic study was undertaken to learn what it was like for a family when a child had IDB. After 36 interviews, the study team was able to create three parent-child dyad personas: archetypes of parents and children with IDB based directly on the data they gathered. These personas were used by the design team to think about how different types of parents and children adapted to the disease and to think what tools should be developed to help different types of parents and children with IDB. An ethnographic study is the method of choice when the goal is to understand a culture, and to present, or explain, its spoken and unspoken nature to people who are not part of the culture, as in the example above of IDB. Before “outsiders” could think about the needs of people with IDB, it was necessary to learn what it is like to live with this disease.

Determining the sample in ethnographic studies typically means using what is called a purposive sample ( Newfield, Sells, Smith, Newfield, & Newfield, 1996 ). Purposive samples are based on criteria that the investigator establishes at the outset, which describe participant characteristics. In the aforementioned IDB example, the criteria were: (1) being a person with IDB who was between 12 and 22 years old or the parent of such a child; (2) being or having a child whose IDB care was provided at one of a particular group of treatment centers; (3) being a pediatric gastrointestinal nurse at one of the centers; or (4) a physician/researcher at one of five treatment centers. Having a sample that is representative of the larger population, always the goal in quantitative research, is not the point in ethnographic studies. Here, the goal is to recruit participants who have the experience to respond to the questions. Out of their intimate knowledge of their culture, the investigator can build a theory, or conceptual model, which could later be tested for generalizability in an entire population.

Ethnographic study designs typically involve a combination of data collection methods. Whenever possible, observing the participants in the midst of whatever experience is the study's focus is desirable. In the process of an ethnographic project on CF, for instance, two students spent a twelve-hour period at the home of a family with a child who had CF, taking notes about what they saw and heard. Interviews with participants are frequently employed to learn more about the experience of interest. An example of this is the work of Sobo and colleagues, who interviewed parents of pediatric patients in a clinic to ask about the barriers they experienced obtaining health care for their child ( Seid, Sobo, Gelhard, & Varni, 2004 ). Diaries and journals detailing people's lived experience may also be used, alone, or in combination with other methods.

Analysis of ethnographic data is variable, depending on the study's goal. One common analytic approach is to begin analysis after the first few interviews have been completed, and to read them to get a sense of their content. The next step is to name the seemingly important words or phrases. At this point, one might begin to see how the names relate to each other; this is the beginning of theory development. This process continues until all the data are collected. At that point, the data are sorted by the names, with data from multiple participants clustered under each topic name ( Boyle, 1994 ). Similar names may be grouped together, or placed under a larger label name (i.e., category). In a sense, what happens is that each interviewer's voice is broken into individual fragments, and everyone's fragments that have the same name are put together. From individual voices speaking on multiple topics, there is now one topic with multiple voices speaking to it.

GROUNDED THEORY RESEARCH

Grounded theory is “grounded” in its data; this inductive approach collects data while simultaneously analyzing it and using the emerging theory to inform data collection ( Rafuls & Moon, 1996 ). This cycle continues until the categories are said to be “saturated,” which typically means the point when no new information is being learned ( Morse, 1995 ). This methodology is generally credited to Glaser and Strauss, who wanted to create a means of developing theoretical models from empirical data ( Charmaz, 2005 ). Perhaps, more than in any other qualitative methodology, the person of the investigator is the key. The extent to which the investigator notices subtle nuances in the data and responds to them with new questions for future participants, or revises an emerging theory, is the extent to which a grounded theory research truly presents a theory capturing the fullness of the data from which it was built. It is also the extent to which the theory is capable of being used to guide future research or alter clinical practice. Grounded theory is the method of choice when there is no existing hypothesis to test. For instance, there was no published data on how parents use faith to cope after their child's diagnosis with CF. Using grounded theory allowed us to develop a theory, or a conceptual model, of how parents used faith to cope ( Grossoehme, Ragsdale, Wooldridge, Cotton, & Seid, 2010 ). An excellent discussion of this method is provided by Charmaz (2006) .

The nature of the research question should dictate the sample description, which should be defined before beginning the data collection. In some cases, the incidence of the phenomena may set some limits on the sample. For example, a study of religious coping by adults who were diagnosed with CF after age 18 years began with a low incidence: this question immediately limited the number of eligible adults in a four-state area to approximately 25 ( Grossoehme et al., 2012 ). Knowing that between 12 and 20 participants might be required in order to have sufficient data to convince ourselves that our categories were indeed saturated, limiting our sample in other ways: for example, selecting representative individuals spread across the number of years since diagnosis would not have made sense. In some studies, the goal is to learn what makes a particular subset of a larger sample special; these subsets are known as “positive deviants” ( Bradley et al., 2009 ).

Once the sample is defined and data collection begins, the analytic process begins shortly thereafter. As will be described in the following paragraphs, interviews and other forms of spoken communication are nearly always transcribed, typically verbatim. Unlike most other qualitative methods, grounded theory uses an iterative design. Sometime around the third or fourth interview has been completed and transcribed and before proceeding with further interviews, it is time to begin analyzing the transcripts. There are two aspects to this. The first is to code the data that you have. Grounded theory prefers to use the participants' own words as the code, rather than having the investigator name it. For example, in the following transcript excerpt, we coded part of the following except:

  • INTERVIEWER: OK. Have your beliefs or perhaps relationship with God changed at all because of what you've gone through the last nine and 10 months with N.?
  • INTERVIEWEE: Yeah, I mean, I feel that I'm stronger than I was before actually.
  • INTERVIEWER: Hmm-hmm. How so? Can you put that into words? I know some of these could be hard to talk about but …
  • INTERVIEWEE: I don't know, I feel like I'm putting his life more in God's hands than I ever was before.

We labeled, or coded, these data as, “I'm putting his life more in God's hands,” whereas in a different methodology we might have simply named it “Trusting God.” Focus on the action in the narrative. Although it can be difficult, you as a researcher must try very hard to set your own ideas aside. Remember you are doing this because there is no pre-existing theory about what you are studying, so you should not be guided by a theory you have in your mind. You must let the data speak for themselves.

The second point is to reflect on the codes and what they are already telling you. What questions are eliciting the narrative data you want? Which ones are not? Questions that are not leading you to the data you want probably need to be changed. Interesting, novel ideas may emerge from the data, or topics that you want to know more about that you did not anticipate and so the interviewer did not' follow up on them. What are the data not telling you that you are seeking? All of this information flows back to revising the semi-structured interview guide ( Charmaz, 2006 ). This issue raised mild concern with the IRB reviewer who had not encountered this methodology before. This concern was overcome by showing that this is an accepted method with voluminous literature behind it, and by showing that the types of item revisions were not expected to significantly alter the study's effect on the participants. From this point onward you collect data, code it, and analyze it simultaneously. As you code a new transcript and come across a statement similar to others, you can begin to put them together. If you are using qualitative analysis software such as NVivo ( “NVivo qualitative data analysis software,” 2012 ), you can make these new codes “children” of a “parent” node (the first statement you encountered on this topic). The next step is called “focused coding” and in this phase you combine what seems to you to be the most significant codes ( Charmaz, 2006 ). These may also be the most frequently occurring, or the topic with the most duplicates, but not necessarily. This is not a quantitative approach in which having large amounts of data is important. You combine codes at this stage in such a way that your new, larger, categories begin to give shape to aspects of the theory you think is going to emerge. As you collect and code more data, and revise your categories, your idea of the theory will change.

Axial coding follows, as you look at your emerging themes or categories, and begin to associate coded data that explains that category. Axial coding refers to coding the words or quotations that are around the category's “axis,” or core. For example, in a study of parental faith and coping in the first year after their child's diagnosis with CF ( Grossoehme et al., 2010 ), one of the categories which emerged was, “Our beliefs have changed.” There were five axial codes which explain aspects of this category. The axial codes were, “Unchanged,” “We've learned how fragile life is,” “Our faith has been strengthened,” “We've gotten away from our parents' viewpoints,” and “I'm better in tune with who I am.” Each of these axial codes had multiple explanatory phrases or sentences under them; together they explain the breadth and dimensions of the category, “Our beliefs have changed.”

The next step is theoretical coding, and here the categories generated during focused coding are synthesized into a theory. Some grounded theorists, notably one of the two most associated with it (Glaser), do not use axial coding but proceed directly to this step as the means of creating coherence out of the data ( Charmaz, 2006 ). As your emerging theory crystallizes, you may pause to see if it has similarities with other theoretical constructs you encountered in your literature search. Does your emerging theory remind you of anything? It would be appropriate to engage in member-checking at this point. In this phase, you show your theoretical model and its supporting categories to participants and ask for their feedback. Does your model make sense to them? Does it help them see this aspect of their experience differently ( Charmaz, 2005 ). Use their feedback to revise your theory and put it in its final form. At this point, you have generated new knowledge: a theory no one has put forth previously, and one that is ready to be tested.

PHENOMENOLOGY RESEARCH

Perhaps the most chaplain-friendly qualitative research approach is phenomenology, because it is all about the search for meaning. Its roots are in the philosophical work of Husserl, Heidegger and Ricoeur ( Boss, Dahl, & Kaplan, 1996 ; Swinton & Mowat, 2006 ). This approach is based on several assumptions: (1) meaning and knowing are social constructions, always incomplete and developing; (2) the investigator is a part of the experience being studied and the investigator's values play a role in the investigation; (3) bias is inherent in all research and should be articulated at the beginning; (4) participants and investigators share knowledge and are partners; (5) common forms of expression (e.g., words or art) are important; and (6) meanings may not be shared by everyone (Boss et al.). John Swinton and Harriet Mowat (2006) described the process of carrying out a phenomenological study of depression and spirituality in adults and reading their book is an excellent way to gain a sense of the whole process. Phenomenology may be the method of choice when you want to study what an experience means to a particular group of people. May not be the best choice when you want to be able to generalize your findings. An accurate presentation of the experience under study is more important in this approach than the ability to claim that the findings apply to across situations or people (Boss et al.). A study of the devil among predominately Hispanic horse track workers is unlikely to be generalizable to experiences of the devil among persons of Scandinavian descent living in Minnesota. Care must be taken not to overstate the findings from a study and extend the conclusions beyond what the data support.

The emphasis on accurately portraying the phenomenon means that large numbers of participants are not required. In fact, relatively small sample sizes are required compared to most quantitative, clinical studies. The goal is to gather descriptions of their lived experience which are rich in detail and imagery, as well as reflection on its theological or psychological meaning. The likelihood of achieving this goal can be enhanced by using a purposeful sample. That is, decide in the beginning approximately how large and how diverse your sample needs to be. For example, CF can be caused by over 1,000 different genetic mutations; some cause more pulmonary symptoms while others cause more gastrointestinal problems. Some people with CF have diabetes and others do not; some have a functioning pancreas and others need to take replacement enzymes before eating or drinking anything other than water. Some CF adolescents may have lung function that is over 100% of what is expected for healthy adolescents of their age and gender, whereas others, with severe pulmonary disease, may have lung function that is just 30% of what is expected for their age and gender. A study of what it is like for an adolescent to live with a life-shortening genetic disease using this approach might benefit from purposive sampling. For example, lung disease severity in CF is broadly described as mild, moderate or severe. A purposeful sample might call for 18 participants divided into 3 age groups (11–13 years; 14–16 years; and 17–19 years old) and disease severity (mild, moderate, and severe). In each of those nine groups there would be one male and one female. In actual practice, one might want to have more than 18 to allow for attrition, but this breakdown gives the basic idea of defining a purposive sample. One could reasonably expect that having the experience of both genders across the spectrum of disease severity and the developmental range of adolescence would permit an accurate, multi-dimensional understanding to emerge of what living with this life-shortening disease means to adolescents. In fact, such an accurate description is more likely to emerge with this purposeful sample of 18 adolescents than with a convenience sample of the first 18 adolescents who might agree to participate in the study during their outpatient clinic appointment. Defining the sample to be studied requires some forethought about what is likely to be needed to gain the fullest understanding of the topic.

Any research design may be used. The design will be dictated by what data are required to understand the phenomena and its meaning. Interviews are by far the most common means of gathering data, although one might also use written texts, such as prayers written in open prayer books in hospital chapels, for example ( ap Sion, 2013 ; Grossoehme, 1996 ), or drawings ( Pendleton, Cavalli, Pargament, & Nasr, 2002 ), or photographs/videos ( Olausson, Ekebergh, & Lindahl, 2012 ). Although the word “text” appears, it should be with the understanding that any form of data is implied.

The theoretical underpinnings of phenomenology, which are beyond the scope of this article, suggest to users that “a method” is unnecessary or indeed, contrary, to phenomenology. However, one phenomenological researcher did articulate a method ( Giorgi, 1985 ), which consists of the following steps. First, the research team immerses themselves in the data. They do this by reading and re-reading the transcribed interviews and listening to the recorded interviews so that they can hear the tone and timbre of the voices. The goal at this stage is to get a sense of the whole. Second, the texts are coded, in which the words, phrases or sentences that stand out as describing the experience or phenomena under study, or which express outright its meaning for the participant are extracted or highlighted. Each coded bit of data is sometimes referred to as a “meaning unit.” Third, similar meaning units are placed into categories. Fourth, for each meaning unit the meaning of the participants' own words is spelled out. For chaplains, this may mean articulating what the experience means in theological language. Other disciplines might transform the participants' words into psychological, sociological or anthropological language. Here the investigators infer the meaning behind the participants' words and articulate it. Finally, each of the transformed statements of meaning are combined into a few thematic statements that describe the experience ( Bassett, 2004 ; Boss et al., 1996 ). After this, it would be appropriate to do member-checking and a subsequent revision of the final model based on participants' responses and feedback.

PRACTICAL CONCERNS

Just as questionnaires or blood samples contain data, in qualitative research it is the recording of people's words, whether in an audio, video, or paper format which hold the data. Interviews, either in-person or by telephone should be recorded using audio, video or both. It is important to have a device with suitable audio quality and fresh batteries. Experience has shown me the benefit of using two audio recorders so that you do not lose data if one of them fails. There are several small recorders available that have USB connections that allow the audio file to be uploaded to a computer easily. To protect participants' privacy, all data should be anonymized by removing any information that could identify individuals. The Standard Operating Procedure in my research group is to replace all participants' names with an “ N .” During the transcription process, all other individuals are identified by their role in square brackets, “[parent].” Depending on the study's goal and the analytic method you have selected, you may want to include symbols for pauses before participants respond, or non-fluencies (e.g., “ummm. …”, “well … uh …”) or non-verbal gestures (if you are video recording). Decide before beginning whether it is important to capture these as data or not. There are conventional symbols which are inserted into transcriptions which capture these data for you. After the initial transcription, these need to be verified by comparing the written copy against the original recording. Verification should be done by someone other than the transcriptionist. There are several tasks at this stage. Depending on the quality of your recording, the clarity of participants' speech and other factors, some words or phrases may have been unintelligible to the transcriptionist, and this is the time to address them. In my research group our Standard Operating Procedure is to highlight unintelligible text during the transcription phase, and a “verifier” attempts three times to clarify the words on the original recording before leaving them marked “unintelligible” in the transcript. No transcriptionist is perfect and if they are unfamiliar with the topic, they may transcribe the recording inaccurately. I recently verified a transcript where a commercial medical transcriptionist changed the participant's gender from “he” to “she” when the word prior to the pronoun ended with an “s.” If this pattern had not been caught during the verification process, it would have been very difficult during the coding to know whether the pronoun referred to the participant or to their daughter.

ETHICAL ISSUES IN QUALITATIVE RESEARCH

Study design.

The issue of power and the possibility of subtle coercion is the concern here. There is an inherent power differential between a research participant and the investigator, which is exacerbated when the investigator is a chaplain. Despite our attempts to be non-threatening, the very words, "chaplain," or "clergy" connote power. For this reason, the chaplain-investigator should not approach potential participants regarding a study. Potential participants may be informed regarding their eligibility to participate by their physician or a chaplain, but the recruitment and informed consent process should be handled by someone else, perhaps a clinical research coordinator. However, as the chaplain-investigator, you will need to teach them how to talk with potential participants about your study and answer their questions. Choose a data collection method that is best-suited to the level of sensitivity of your research topic. Focus groups can provide data with multiple perspectives, and they are a poor choice when there may be pressure to provide socially correct responses, or when disclosures may be stigmatizing. In such cases, it is better to collect data using individual semi-structured interviews.

Develop a plan for assessing participants' discomfort, anxiety, or even more severe reactions during the study. For instance, what will you do when someone discloses his/her current thoughts of self-harm, or experiences a flashback to a prior traumatic event that was triggered during an interview? How will you handle this if you are collecting data in person? By telephone? You will need to be specific who must be informed and who will make decisions about responding to the risk.

Privacy and Confidentiality

In addition to maintaining privacy and confidentiality of your actual data and other study documents, consider how you will protect participants' privacy when you write the study up for publication. Make sure that people cannot be identified by their quotations that you include as you publish data. The smaller the population you are working with, the more diligently you need to work on this. If the transcriptionist is not an employee of your institution and under the same privacy and confidentiality policies, it is up to you to ensure that an external transcriptionist takes steps to protect and maintain the privacy of participants' data.

Qualitative research is an accessible way for chaplains to contribute new knowledge regarding the sacred dimension of people's lived experience. Chaplains are already sensitive to and familiar with many aspects of qualitative research methodologies. Studies need to be designed to be valid and meaningful, and are best done collaboratively. They provide an excellent opportunity to develop working relationships with physicians, medical anthropologists, nurses, psychologists, and sociologists, all of whom have rich traditions of qualitative research. This article can only provide an overview of some of the issues related to qualitative research and some of its methods. The texts cited, as well as others, provide additional information needed before designing and carrying out a qualitative study. Qualitative research is a tool that chaplains can use to develop new knowledge and contribute to professional chaplaincy's ability to facilitate the healing of brokenness and disease.

Publisher's Disclaimer: Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and should be independently verified with primary sources of information. Taylor and Francis shall not be liable for any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of the Content.

  • ap Sion T. Coping through prayer: An empirical study in implicit religion concerning prayers for children in hospital. Mental Health, Religion & Culture. 2013; 16 (9):936–952. doi: 10.1080/13674676.2012.756186. [ Google Scholar ]
  • Bassett C. Phenomenology. In: Bassett C, editor. Qualitative research in health care. Whurr Publishers, Ltd; London, UK: 2004. pp. 154–177. [ Google Scholar ]
  • Boss P, Dahl C, Kaplan L. The use of phenomenology for family therapy research. In: Sprenkle DH, Moon SM, editors. Research methods in family therapy. Guilford Press; New York, NY: 1996. pp. 83–106. [ Google Scholar ]
  • Boyle J. Styles of ethnography. In: Morse JM, editor. Critical issues in qualitative research methods. Sage Publications; Thousand Oaks, CA: 1994. pp. 159–185. [ Google Scholar ]
  • Bradley EH, Curry LA, Ramanadhan S, Rowe L, Nembhard IM, Krumholz HM. Research in action: Using positive deviance to improve quality of health care. Implementation Science. 2009; 4 (25) doi: 10.1186/1748-5908-4-25. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Charmaz K. Grounded theory in the 21st century. In: Denzin NK, Lincoln YS, editors. The SAGE handbook of qualitative research. Sage Publications; Thousand Oaks, CA: 2005. pp. 507–535. [ Google Scholar ]
  • Charmaz K. Constructing grounded theory. Sage Publications; Thousand Oaks, CA: 2006. [ Google Scholar ]
  • Cohen D, Crabtree B. Qualitative research guidelines project. 2006 Retrieved March 11, 2014, from http://www.qualres.org/HomeSemi-3629.html .
  • Fore D, Goldenhar LM, Margolis PA, Seid M. Using goal-directed design to create a novel system for improving chronic illness. JMIR Research Protocols. 2013; 2 (2):343. doi: 10.2196/resprot.2749. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Giorgi A. Phenomenology and psychological research. Dusquesne University Press; Pittsburgh, PA: 1985. [ Google Scholar ]
  • Grossoehme DH. Prayer reveals belief: Images of God from hospital prayers. Journal of Pastoral Care. 1996; 50 (1):33–39. [ PubMed ] [ Google Scholar ]
  • Grossoehme DH. Research as a chaplaincy intervention. Journal of Health Care Chaplaincy. 2011; 17 (3–4):97–99. doi: 10.1080/08854726.2011.616165. [ PubMed ] [ Google Scholar ]
  • Grossoehme DH, Cotton S, Ragsdale J, Quittner AL, McPhail G, Seid M. "I honestly believe God keeps me healthy so I can take care of my child": Parental use of faith related to treatment adherence. Journal of Health Care Chaplaincy. 2013; 19 (2):66–78. doi: 10.1080/08854726.2013.779540. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Grossoehme DH, Ragsdale JR, Cotton S, Meyers MA, Clancy JP, Seid M, Joseph PM. Using spirituality after an adult CF diagnosis: Cognitive reframing and adherence motivation. Journal of Health Care Chaplaincy. 2012; 18 (3–4):110–120. doi: 10.1080/08854726.2012.720544. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Grossoehme DH, Ragsdale J, Wooldridge JL, Cotton S, Seid M. We can handle this: Parents' use of religion in the first year following their child's diagnosis with cystic fibrosis. Journal of Health Care Chaplaincy. 2010; 16 (3–4):95–108. doi: 10.1080/08854726.2010.480833. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Malterud K. Qualitative research: Standards, challenges, and guidelines. The Lancet. 2001; 358 (9280):483–488. [ PubMed ] [ Google Scholar ]
  • Morse JM. The significance of saturation. Qualitative Health Research. 1995; 5 :147–149. [ Google Scholar ]
  • Newfield N, Sells SP, Smith TE, Newfield S, Newfield F. Ethnographic research methods. In: Sprenkle DH, Moon SM, editors. Research methods in family therapy. The Guilford Press; New York, NY: 1996. pp. 25–63. [ Google Scholar ]
  • NVivo qualitative data analysis software . QSR International Pty Ltd. 2012. [ Google Scholar ]
  • Olausson S, Ekebergh M, Lindahl B. The ICU patient room: Views and meanings as experienced by the next of kin: A phenomenological hermeneutical study. Intensive and Critical Care Nursing. 2012; 28 (3):176–184. [ PubMed ] [ Google Scholar ]
  • Pendleton SM, Cavalli KS, Pargament KI, Nasr SZ. Religious/spiritual coping in childhood cystic fibrosis: A qualitative study. Pediatrics. 2002; 109 (1):E8. [ PubMed ] [ Google Scholar ]
  • Rafuls SE, Moon SM. Grounded theory methodology in family therapy research. In: Sprenkle DH, Moon SM, editors. Research methods in family therapy. The Guilford Press; New York, NY: 1996. [ Google Scholar ]
  • Risk JL. Building a new life: A chaplain's theory based case study of chronic illness. Journal of Health Care Chaplaincy. 2013; 19 :81–98. [ PubMed ] [ Google Scholar ]
  • Seid M, Sobo EJ, Gelhard LR, Varni JW. Parents' reports of barriers to care for children with special health care needs: Development and validation of the barriers to care questionnaire. Ambulatory Pediatrics. 2004; 4 (4):323–331. doi: 10.1367/A03-198R.1. [ PubMed ] [ Google Scholar ]
  • Sobo EJ. Culture and meaning in health services research. Left Coast Press, Inc; Walnut Creek, CA: 2009. [ Google Scholar ]
  • Swinton J, Mowat H. Practical theology and qualitative research. SCM Press; London, UK: 2006. [ Google Scholar ]
  • Tong A, Sainsbury P, Craig J. Consolidated criteria for reporting qualitative research (COREQ): A 32-item checklist for interviews and focus groups. International Journal for Quality in Health Care. 2007; 19 (6):349–357. [ PubMed ] [ Google Scholar ]

IMAGES

  1. Methodological matrix of organizational ethnographic case studies for

    ethnographic case study methodology

  2. Case study and Ethnography

    ethnographic case study methodology

  3. PPT

    ethnographic case study methodology

  4. Ethnographic research method

    ethnographic case study methodology

  5. What is Ethnographic Research Method?

    ethnographic case study methodology

  6. Case study and Ethnography

    ethnographic case study methodology

VIDEO

  1. Webinar: Case study methodology for psychotherapy process research

  2. Ethnographic Study Tour at SDA Banting

  3. History of Ethnobotany

  4. Episode 27: IIDS Webinar on Politics, Economy and Public Policy

  5. Ethnographic Research Methodology

  6. research|qualitative research design|types of qualitative research design|ethnographic|grounded

COMMENTS

  1. Organizational Ethnographic Case Studies: Toward a New Generative In

    While previous research has focused separately on the methodological legitimacy of new forms of organizational ethnography and case study designs applied to health care research, little is known about why and when it is better to combine these two approaches (Baxter & Jack, 2008; Bryman & Buchanan, 2018; Carolan et al., 2016; Fine et al., 2009; Hammersley, 2018; Lambert et al., 2011; Ybema et ...

  2. Practices of Ethnographic Research: Introduction to the Special Issue

    Methods and practices of ethnographic research are closely connected: practices inform methods, and methods inform practices. In a recent study on the history of qualitative research, Ploder (2018) found that methods are typically developed by researchers conducting pioneering studies that deal with an unknown phenomenon or field (a study of Andreas Franzmann 2016 points in a similar direction).

  3. Focused ethnographic case studies, methodology and description of sites

    This study used team-focused ethnographic methods. In a focused ethnography, rather than embedding a single researcher in a social setting for a lengthy period, more targeted data collection is used to explore the study topics. Using existing information from the literature and from what is known in clinical practice helps to determine the research question and, subsequently, to generate ...

  4. Ethnographic Case Studies

    The author combined a single-case study approach with ethnographic methods to "engage in close analysis of classroom language use and the discursive negotiation of identities and ideologies, while situating these analyses within a rich understanding of the sociolinguistic context of this TWI classroom" (p. 78-79). She employed various ...

  5. PDF Comparing the Five Approaches

    an ethnography. An in-depth study of a bounded system or a case (or several cases) becomes a case study. The general structures of the written report may be used in designing a journal-article-length study. However, because of the numerous steps in each, they also have applicability as chapters of a dissertation or a book-length work.

  6. What Is Ethnography?

    Ethnography is a sensitive research method, and it may take multiple attempts to find a feasible approach. Working with informants. ... Examples & Methods A case study is a detailed study of a specific subject in its real-world context. It can focus on a person, group, event or organization. 3846.

  7. Blending the Focused Ethnographic Method and Case Study Research

    In this article, we present the benefits of blending the methodological characteristics of the focused ethnographic method (FEM) and case study research (CSR) for a study on auxiliary work processes in a hospital in Barcelona, Spain. We argue that incorporating CSR logic and principles in the FEM produces a better form of inquiry, as this ...

  8. Qualitative research methodologies: ethnography

    The previous articles (there were 2 before this 1) in this series discussed several methodological approaches commonly used by qualitative researchers in the health professions. This article focuses on another important qualitative methodology: ethnography. It provides background for those who will encounter this methodology in their reading rather than instructions for carrying out such research.

  9. Application of case study research and ethnography methods: Lessons

    Case study research is often used to gain a deeper understanding of patient care issues. •. A qualitatively-driven mixed-method design can be used within healthcare research. •. Case study designs can be enhanced by using focused ethnography to examine cultural content. •.

  10. PDF ETHNOGRAPHIC RESEARCH

    • How to write and evaluate ethnographic research. WHAT IS ETHNOGRAPHIC RESEARCH? Ethnographic research takes a cultural lens to the study of people's lives within their communities (Hammersley and Atkinson, 2007; Fetterman, 2010). The roots of ethno­ graphy lie in anthropological studies that focused on studying social and cultural

  11. PDF International Journal of Education & Literacy Studies

    Ethnographic and Case Study Approaches: Philosophical and Methodological Analysis 151 characteristics of both ethnography and case study then will ... a typical case, using a number of methods to prevent errors and distortions. Philosophical and Methodological Perspectives We now have insights into the definition, purpose and char- ...

  12. An Ethnographic Case Study Design

    Although an ethnographic case study design appears to be a qualitative approach, it is in practice a mixed-method approach including elements of both qualitative and quantitative methods (Yin, 2014) for triangulation purposes ( Holloway, Brown, & Shipway, 2010 ). Thus, such an approach may be a preferred strategy to answer how, what, or why ...

  13. Two cases of ethnography: Grounded theory and the extended case method

    University of California, Los Angeles, USA. ABSTRACT« Sociological ethnography largely draws upon two epistemologically competing perspectives - grounded theory and the extended case method - with a different conceptualization of sociological case-construction and theory. We argue that the sociological case in the extended case method is ...

  14. (PDF) Comparing Case Study and Ethnography as ...

    Selecting a case study as the design also came with the benefit that a case study can "follow ethnographic methods" in describing a case whereas "ethnographers do not always produce case studies ...

  15. Organizational Ethnographic Case Studies: Toward a New Generative In

    organizational ethnographic case studies have their own distinct methodological identity in the wider domain of qualitative health ... is an ethnographic-adapted method that study social phenomena that cannot be explained by focusing on a single site (Bikker et al., 2017; Marcus, 2011; 2012; Molloy et al., 2017).

  16. Sage Research Methods Cases Part 1

    Abstract. This article presents my experience for conducting an ethnographic case study entitled Building Strong Bridges between the Museum and Its Community: An Ethnographic Understanding of the Culture and Systems of One Community's Art Museum.In this article, I focus on the complex processes, flexible nature, and unexpected encounters of qualitative research by sharing my experiences and ...

  17. Ethnographic research as an evolving method for supporting healthcare

    The relationship between ethnography and healthcare improvement has been the subject of methodological concern. We conducted a scoping review of ethnographic literature on healthcare improvement topics, with two aims: (1) to describe current ethnographic methods and practices in healthcare improvement research and (2) to consider how these may affect habit and skill formation in the service of ...

  18. Use ethnographic methods & participant observation

    Participant observation is one of the main ethnographic data collection methods. The essence of participant observation is that you, as the researcher, observe the subject of research, either by participating directly in the action, as a member of the study population, or as a "pure" observer, in which case you do not participate in the action ...

  19. Ethnographic Research -Types, Methods and Guide

    An ethnographic study is a research method that involves the detailed and systematic study of a particular group, culture, or community. Ethnographic studies seek to understand the beliefs, values, behaviors, and social dynamics of a group through direct observation and participation in their daily life. ... Case Study - Methods, Examples and ...

  20. Ethnographic research as an evolving method for supporting healthcare

    Information about study aims, methodology and recommendations for improvement were extracted. ... Leslie M, Minion J, Martin GP, Coleman JJ. Improving quality and safety of care using "technovigilance": an ethnographic case study of secondary use of data from an electronic prescribing and decision support system. Milbank Q. 2013; 91 (3):424 ...

  21. Case Study Methodology of Qualitative Research: Key Attributes and

    A case study is one of the most commonly used methodologies of social research. This article attempts to look into the various dimensions of a case study research strategy, the different epistemological strands which determine the particular case study type and approach adopted in the field, discusses the factors which can enhance the effectiveness of a case study research, and the debate ...

  22. Research MethodologyOverview of Qualitative Research

    An ethnographic study is the method of choice when the goal is to understand a culture, and to present, or explain, its spoken and unspoken nature to people who are not part of the culture, as in the example above of IDB. ... Building a new life: A chaplain's theory based case study of chronic illness. Journal of Health Care Chaplaincy. 2013 ...

  23. Is Microethnography an Ethnographic Case Study? and/or a mini

    This points to what separates mini-ethnography case study from ethnographic case study. Compared to ethnographic case study, mini-ethnography case study is more appropriate for researchers who have very short time to spend in the field, but who still want to have an in-depth understanding of the phenomenon understudy (Dooley et al., 2020).