Qualitative Research : Definition

Qualitative research is the naturalistic study of social meanings and processes, using interviews, observations, and the analysis of texts and images.  In contrast to quantitative researchers, whose statistical methods enable broad generalizations about populations (for example, comparisons of the percentages of U.S. demographic groups who vote in particular ways), qualitative researchers use in-depth studies of the social world to analyze how and why groups think and act in particular ways (for instance, case studies of the experiences that shape political views).   

Events and Workshops

  • Introduction to NVivo Have you just collected your data and wondered what to do next? Come join us for an introductory session on utilizing NVivo to support your analytical process. This session will only cover features of the software and how to import your records. Please feel free to attend any of the following sessions below: April 25th, 2024 12:30 pm - 1:45 pm Green Library - SVA Conference Room 125 May 9th, 2024 12:30 pm - 1:45 pm Green Library - SVA Conference Room 125
  • Next: Choose an approach >>
  • Choose an approach
  • Find studies
  • Learn methods
  • Getting Started
  • Get software
  • Get data for secondary analysis
  • Network with researchers

Profile Photo

  • Last Updated: Aug 9, 2024 2:09 PM
  • URL: https://guides.library.stanford.edu/qualitative_research

Have a language expert improve your writing

Run a free plagiarism check in 10 minutes, automatically generate references for free.

  • Knowledge Base
  • Methodology
  • What Is Qualitative Research? | Methods & Examples

What Is Qualitative Research? | Methods & Examples

Published on 4 April 2022 by Pritha Bhandari . Revised on 30 January 2023.

Qualitative research involves collecting and analysing non-numerical data (e.g., text, video, or audio) to understand concepts, opinions, or experiences. It can be used to gather in-depth insights into a problem or generate new ideas for research.

Qualitative research is the opposite of quantitative research , which involves collecting and analysing numerical data for statistical analysis.

Qualitative research is commonly used in the humanities and social sciences, in subjects such as anthropology, sociology, education, health sciences, and history.

  • How does social media shape body image in teenagers?
  • How do children and adults interpret healthy eating in the UK?
  • What factors influence employee retention in a large organisation?
  • How is anxiety experienced around the world?
  • How can teachers integrate social issues into science curriculums?

Table of contents

Approaches to qualitative research, qualitative research methods, qualitative data analysis, advantages of qualitative research, disadvantages of qualitative research, frequently asked questions about qualitative research.

Qualitative research is used to understand how people experience the world. While there are many approaches to qualitative research, they tend to be flexible and focus on retaining rich meaning when interpreting data.

Common approaches include grounded theory, ethnography, action research, phenomenological research, and narrative research. They share some similarities, but emphasise different aims and perspectives.

Qualitative research approaches
Approach What does it involve?
Grounded theory Researchers collect rich data on a topic of interest and develop theories .
Researchers immerse themselves in groups or organisations to understand their cultures.
Researchers and participants collaboratively link theory to practice to drive social change.
Phenomenological research Researchers investigate a phenomenon or event by describing and interpreting participants’ lived experiences.
Narrative research Researchers examine how stories are told to understand how participants perceive and make sense of their experiences.

Prevent plagiarism, run a free check.

Each of the research approaches involve using one or more data collection methods . These are some of the most common qualitative methods:

  • Observations: recording what you have seen, heard, or encountered in detailed field notes.
  • Interviews:  personally asking people questions in one-on-one conversations.
  • Focus groups: asking questions and generating discussion among a group of people.
  • Surveys : distributing questionnaires with open-ended questions.
  • Secondary research: collecting existing data in the form of texts, images, audio or video recordings, etc.
  • You take field notes with observations and reflect on your own experiences of the company culture.
  • You distribute open-ended surveys to employees across all the company’s offices by email to find out if the culture varies across locations.
  • You conduct in-depth interviews with employees in your office to learn about their experiences and perspectives in greater detail.

Qualitative researchers often consider themselves ‘instruments’ in research because all observations, interpretations and analyses are filtered through their own personal lens.

For this reason, when writing up your methodology for qualitative research, it’s important to reflect on your approach and to thoroughly explain the choices you made in collecting and analysing the data.

Qualitative data can take the form of texts, photos, videos and audio. For example, you might be working with interview transcripts, survey responses, fieldnotes, or recordings from natural settings.

Most types of qualitative data analysis share the same five steps:

  • Prepare and organise your data. This may mean transcribing interviews or typing up fieldnotes.
  • Review and explore your data. Examine the data for patterns or repeated ideas that emerge.
  • Develop a data coding system. Based on your initial ideas, establish a set of codes that you can apply to categorise your data.
  • Assign codes to the data. For example, in qualitative survey analysis, this may mean going through each participant’s responses and tagging them with codes in a spreadsheet. As you go through your data, you can create new codes to add to your system if necessary.
  • Identify recurring themes. Link codes together into cohesive, overarching themes.

There are several specific approaches to analysing qualitative data. Although these methods share similar processes, they emphasise different concepts.

Qualitative data analysis
Approach When to use Example
To describe and categorise common words, phrases, and ideas in qualitative data. A market researcher could perform content analysis to find out what kind of language is used in descriptions of therapeutic apps.
To identify and interpret patterns and themes in qualitative data. A psychologist could apply thematic analysis to travel blogs to explore how tourism shapes self-identity.
To examine the content, structure, and design of texts. A media researcher could use textual analysis to understand how news coverage of celebrities has changed in the past decade.
To study communication and how language is used to achieve effects in specific contexts. A political scientist could use discourse analysis to study how politicians generate trust in election campaigns.

Qualitative research often tries to preserve the voice and perspective of participants and can be adjusted as new research questions arise. Qualitative research is good for:

  • Flexibility

The data collection and analysis process can be adapted as new ideas or patterns emerge. They are not rigidly decided beforehand.

  • Natural settings

Data collection occurs in real-world contexts or in naturalistic ways.

  • Meaningful insights

Detailed descriptions of people’s experiences, feelings and perceptions can be used in designing, testing or improving systems or products.

  • Generation of new ideas

Open-ended responses mean that researchers can uncover novel problems or opportunities that they wouldn’t have thought of otherwise.

Researchers must consider practical and theoretical limitations in analysing and interpreting their data. Qualitative research suffers from:

  • Unreliability

The real-world setting often makes qualitative research unreliable because of uncontrolled factors that affect the data.

  • Subjectivity

Due to the researcher’s primary role in analysing and interpreting data, qualitative research cannot be replicated . The researcher decides what is important and what is irrelevant in data analysis, so interpretations of the same data can vary greatly.

  • Limited generalisability

Small samples are often used to gather detailed data about specific contexts. Despite rigorous analysis procedures, it is difficult to draw generalisable conclusions because the data may be biased and unrepresentative of the wider population .

  • Labour-intensive

Although software can be used to manage and record large amounts of text, data analysis often has to be checked or performed manually.

Quantitative research deals with numbers and statistics, while qualitative research deals with words and meanings.

Quantitative methods allow you to test a hypothesis by systematically collecting and analysing data, while qualitative methods allow you to explore ideas and experiences in depth.

There are five common approaches to qualitative research :

  • Grounded theory involves collecting data in order to develop new theories.
  • Ethnography involves immersing yourself in a group or organisation to understand its culture.
  • Narrative research involves interpreting stories to understand how people make sense of their experiences and perceptions.
  • Phenomenological research involves investigating phenomena through people’s lived experiences.
  • Action research links theory and practice in several cycles to drive innovative changes.

Data collection is the systematic process by which observations or measurements are gathered in research. It is used in many different contexts by academics, governments, businesses, and other organisations.

There are various approaches to qualitative data analysis , but they all share five steps in common:

  • Prepare and organise your data.
  • Review and explore your data.
  • Develop a data coding system.
  • Assign codes to the data.
  • Identify recurring themes.

The specifics of each step depend on the focus of the analysis. Some common approaches include textual analysis , thematic analysis , and discourse analysis .

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 Reference Generator.

Bhandari, P. (2023, January 30). What Is Qualitative Research? | Methods & Examples. Scribbr. Retrieved 16 September 2024, from https://www.scribbr.co.uk/research-methods/introduction-to-qualitative-research/

Is this article helpful?

Pritha Bhandari

Pritha Bhandari

Educational resources and simple solutions for your research journey

What is qualitative research? Methods, types, approaches, and examples

What is Qualitative Research? Methods, Types, Approaches and Examples

Qualitative research is a type of method that researchers use depending on their study requirements. Research can be conducted using several methods, but before starting the process, researchers should understand the different methods available to decide the best one for their study type. The type of research method needed depends on a few important criteria, such as the research question, study type, time, costs, data availability, and availability of respondents. The two main types of methods are qualitative research and quantitative research. Sometimes, researchers may find it difficult to decide which type of method is most suitable for their study. Keeping in mind a simple rule of thumb could help you make the correct decision. Quantitative research should be used to validate or test a theory or hypothesis and qualitative research should be used to understand a subject or event or identify reasons for observed patterns.  

Qualitative research methods are based on principles of social sciences from several disciplines like psychology, sociology, and anthropology. In this method, researchers try to understand the feelings and motivation of their respondents, which would have prompted them to select or give a particular response to a question. Here are two qualitative research examples :  

  • Two brands (A & B) of the same medicine are available at a pharmacy. However, Brand A is more popular and has higher sales. In qualitative research , the interviewers would ideally visit a few stores in different areas and ask customers their reason for selecting either brand. Respondents may have different reasons that motivate them to select one brand over the other, such as brand loyalty, cost, feedback from friends, doctor’s suggestion, etc. Once the reasons are known, companies could then address challenges in that specific area to increase their product’s sales.  
  • A company organizes a focus group meeting with a random sample of its product’s consumers to understand their opinion on a new product being launched.  

qualitative meaning of research

Table of Contents

What is qualitative research? 1

Qualitative research is the process of collecting, analyzing, and interpreting non-numerical data. The findings of qualitative research are expressed in words and help in understanding individuals’ subjective perceptions about an event, condition, or subject. This type of research is exploratory and is used to generate hypotheses or theories from data. Qualitative data are usually in the form of text, videos, photographs, and audio recordings. There are multiple qualitative research types , which will be discussed later.  

Qualitative research methods 2

Researchers can choose from several qualitative research methods depending on the study type, research question, the researcher’s role, data to be collected, etc.  

The following table lists the common qualitative research approaches with their purpose and examples, although there may be an overlap between some.  

     
Narrative  Explore the experiences of individuals and tell a story to give insight into human lives and behaviors. Narratives can be obtained from journals, letters, conversations, autobiographies, interviews, etc.  A researcher collecting information to create a biography using old documents, interviews, etc. 
Phenomenology  Explain life experiences or phenomena, focusing on people’s subjective experiences and interpretations of the world.  Researchers exploring the experiences of family members of an individual undergoing a major surgery.  
Grounded theory  Investigate process, actions, and interactions, and based on this grounded or empirical data a theory is developed. Unlike experimental research, this method doesn’t require a hypothesis theory to begin with.  A company with a high attrition rate and no prior data may use this method to understand the reasons for which employees leave. 
Ethnography  Describe an ethnic, cultural, or social group by observation in their naturally occurring environment.  A researcher studying medical personnel in the immediate care division of a hospital to understand the culture and staff behaviors during high capacity. 
Case study  In-depth analysis of complex issues in real-life settings, mostly used in business, law, and policymaking. Learnings from case studies can be implemented in other similar contexts.  A case study about how a particular company turned around its product sales and the marketing strategies they used could help implement similar methods in other companies. 

Types of qualitative research 3,4

The data collection methods in qualitative research are designed to assess and understand the perceptions, motivations, and feelings of the respondents about the subject being studied. The different qualitative research types include the following:  

  • In-depth or one-on-one interviews : This is one of the most common qualitative research methods and helps the interviewers understand a respondent’s subjective opinion and experience pertaining to a specific topic or event. These interviews are usually conversational and encourage the respondents to express their opinions freely. Semi-structured interviews, which have open-ended questions (where the respondents can answer more than just “yes” or “no”), are commonly used. Such interviews can be either face-to-face or telephonic, and the duration can vary depending on the subject or the interviewer. Asking the right questions is essential in this method so that the interview can be led in the suitable direction. Face-to-face interviews also help interviewers observe the respondents’ body language, which could help in confirming whether the responses match.  
  • Document study/Literature review/Record keeping : Researchers’ review of already existing written materials such as archives, annual reports, research articles, guidelines, policy documents, etc.  
  • Focus groups : Usually include a small sample of about 6-10 people and a moderator, to understand the participants’ opinion on a given topic. Focus groups ensure constructive discussions to understand the why, what, and, how about the topic. These group meetings need not always be in-person. In recent times, online meetings are also encouraged, and online surveys could also be administered with the option to “write” subjective answers as well. However, this method is expensive and is mostly used for new products and ideas.  
  • Qualitative observation : In this method, researchers collect data using their five senses—sight, smell, touch, taste, and hearing. This method doesn’t include any measurements but only the subjective observation. For example, “The dessert served at the bakery was creamy with sweet buttercream frosting”; this observation is based on the taste perception.  

qualitative meaning of research

Qualitative research : Data collection and analysis

  • Qualitative data collection is the process by which observations or measurements are gathered in research.  
  • The data collected are usually non-numeric and subjective and could be recorded in various methods, for instance, in case of one-to-one interviews, the responses may be recorded using handwritten notes, and audio and video recordings, depending on the interviewer and the setting or duration.  
  • Once the data are collected, they should be transcribed into meaningful or useful interpretations. An experienced researcher could take about 8-10 hours to transcribe an interview’s recordings. All such notes and recordings should be maintained properly for later reference.  
  • Some interviewers make use of “field notes.” These are not exactly the respondents’ answers but rather some observations the interviewer may have made while asking questions and may include non-verbal cues or any information about the setting or the environment. These notes are usually informal and help verify respondents’ answers.  

2. Qualitative data analysis 

  • This process involves analyzing all the data obtained from the qualitative research methods in the form of text (notes), audio-video recordings, and pictures.  
  • Text analysis is a common form of qualitative data analysis in which researchers examine the social lives of the participants and analyze their words, actions, etc. in specific contexts. Social media platforms are now playing an important role in this method with researchers analyzing all information shared online.   

There are usually five steps in the qualitative data analysis process: 5

  • Prepare and organize the data  
  • Transcribe interviews  
  • Collect and document field notes and other material  
  • Review and explore the data  
  • Examine the data for patterns or important observations  
  • Develop a data coding system  
  • Create codes to categorize and connect the data  
  • Assign these codes to the data or responses  
  • Review the codes  
  • Identify recurring themes, opinions, patterns, etc.  
  • Present the findings  
  • Use the best possible method to present your observations  

The following table 6 lists some common qualitative data analysis methods used by companies to make important decisions, with examples and when to use each. The methods may be similar and can overlap.  

     
Content analysis  To identify patterns in text, by grouping content into words, concepts, and themes; that is, determine presence of certain words or themes in some text  Researchers examining the language used in a journal article to search for bias 
Narrative analysis  To understand people’s perspectives on specific issues. Focuses on people’s stories and the language used to tell these stories  A researcher conducting one or several in-depth interviews with an individual over a long period 
Discourse analysis  To understand political, cultural, and power dynamics in specific contexts; that is, how people express themselves in different social contexts  A researcher studying a politician’s speeches across multiple contexts, such as audience, region, political history, etc. 
Thematic analysis  To interpret the meaning behind the words used by people. This is done by identifying repetitive patterns or themes by reading through a dataset  Researcher analyzing raw data to explore the impact of high-stakes examinations on students and parents 

Characteristics of qualitative research methods 4

  • Unstructured raw data : Qualitative research methods use unstructured, non-numerical data , which are analyzed to generate subjective conclusions about specific subjects, usually presented descriptively, instead of using statistical data.  
  • Site-specific data collection : In qualitative research methods , data are collected at specific areas where the respondents or researchers are either facing a challenge or have a need to explore. The process is conducted in a real-world setting and participants do not need to leave their original geographical setting to be able to participate.  
  • Researchers’ importance : Researchers play an instrumental role because, in qualitative research , communication with respondents is an essential part of data collection and analysis. In addition, researchers need to rely on their own observation and listening skills during an interaction and use and interpret that data appropriately.  
  • Multiple methods : Researchers collect data through various methods, as listed earlier, instead of relying on a single source. Although there may be some overlap between the qualitative research methods , each method has its own significance.  
  • Solving complex issues : These methods help in breaking down complex problems into more useful and interpretable inferences, which can be easily understood by everyone.  
  • Unbiased responses : Qualitative research methods rely on open communication where the participants are allowed to freely express their views. In such cases, the participants trust the interviewer, resulting in unbiased and truthful responses.  
  • Flexible : The qualitative research method can be changed at any stage of the research. The data analysis is not confined to being done at the end of the research but can be done in tandem with data collection. Consequently, based on preliminary analysis and new ideas, researchers have the liberty to change the method to suit their objective.  

qualitative meaning of research

When to use qualitative research   4

The following points will give you an idea about when to use qualitative research .  

  • When the objective of a research study is to understand behaviors and patterns of respondents, then qualitative research is the most suitable method because it gives a clear insight into the reasons for the occurrence of an event.  
  • A few use cases for qualitative research methods include:  
  • New product development or idea generation  
  • Strengthening a product’s marketing strategy  
  • Conducting a SWOT analysis of product or services portfolios to help take important strategic decisions  
  • Understanding purchasing behavior of consumers  
  • Understanding reactions of target market to ad campaigns  
  • Understanding market demographics and conducting competitor analysis  
  • Understanding the effectiveness of a new treatment method in a particular section of society  

A qualitative research method case study to understand when to use qualitative research 7

Context : A high school in the US underwent a turnaround or conservatorship process and consequently experienced a below average teacher retention rate. Researchers conducted qualitative research to understand teachers’ experiences and perceptions of how the turnaround may have influenced the teachers’ morale and how this, in turn, would have affected teachers’ retention.  

Method : Purposive sampling was used to select eight teachers who were employed with the school before the conservatorship process and who were subsequently retained. One-on-one semi-structured interviews were conducted with these teachers. The questions addressed teachers’ perspectives of morale and their views on the conservatorship process.  

Results : The study generated six factors that may have been influencing teachers’ perspectives: powerlessness, excessive visitations, loss of confidence, ineffective instructional practices, stress and burnout, and ineffective professional development opportunities. Based on these factors, four recommendations were made to increase teacher retention by boosting their morale.  

qualitative meaning of research

Advantages of qualitative research 1

  • Reflects real-world settings , and therefore allows for ambiguities in data, as well as the flexibility to change the method based on new developments.  
  • Helps in understanding the feelings or beliefs of the respondents rather than relying only on quantitative data.  
  • Uses a descriptive and narrative style of presentation, which may be easier to understand for people from all backgrounds.  
  • Some topics involving sensitive or controversial content could be difficult to quantify and so qualitative research helps in analyzing such content.  
  • The availability of multiple data sources and research methods helps give a holistic picture.  
  • There’s more involvement of participants, which gives them an assurance that their opinion matters, possibly leading to unbiased responses.   

Disadvantages of qualitative research 1

  • Large-scale data sets cannot be included because of time and cost constraints.  
  • Ensuring validity and reliability may be a challenge because of the subjective nature of the data, so drawing definite conclusions could be difficult.  
  • Replication by other researchers may be difficult for the same contexts or situations.  
  • Generalization to a wider context or to other populations or settings is not possible.  
  • Data collection and analysis may be time consuming.  
  • Researcher’s interpretation may alter the results causing an unintended bias.  

Differences between qualitative research and quantitative research 1

     
Purpose and design  Explore ideas, formulate hypotheses; more subjective  Test theories and hypotheses, discover causal relationships; measurable and more structured 
Data collection method  Semi-structured interviews/surveys with open-ended questions, document study/literature reviews, focus groups, case study research, ethnography  Experiments, controlled observations, questionnaires and surveys with a rating scale or closed-ended questions. The methods can be experimental, quasi-experimental, descriptive, or correlational. 
Data analysis  Content analysis (determine presence of certain words/concepts in texts), grounded theory (hypothesis creation by data collection and analysis), thematic analysis (identify important themes/patterns in data and use these to address an issue)  Statistical analysis using applications such as Excel, SPSS, R 
Sample size  Small  Large 
Example  A company organizing focus groups or one-to-one interviews to understand customers’ (subjective) opinions about a specific product, based on which the company can modify their marketing strategy  Customer satisfaction surveys sent out by companies. Customers are asked to rate their experience on a rating scale of 1 to 5  

Frequently asked questions on qualitative research  

Q: how do i know if qualitative research is appropriate for my study  .

A: Here’s a simple checklist you could use:  

  • Not much is known about the subject being studied.  
  • There is a need to understand or simplify a complex problem or situation.  
  • Participants’ experiences/beliefs/feelings are required for analysis.  
  • There’s no existing hypothesis to begin with, rather a theory would need to be created after analysis.  
  • You need to gather in-depth understanding of an event or subject, which may not need to be supported by numeric data.  

Q: How do I ensure the reliability and validity of my qualitative research findings?  

A: To ensure the validity of your qualitative research findings you should explicitly state your objective and describe clearly why you have interpreted the data in a particular way. Another method could be to connect your data in different ways or from different perspectives to see if you reach a similar, unbiased conclusion.   

To ensure reliability, always create an audit trail of your qualitative research by describing your steps and reasons for every interpretation, so that if required, another researcher could trace your steps to corroborate your (or their own) findings. In addition, always look for patterns or consistencies in the data collected through different methods.  

Q: Are there any sampling strategies or techniques for qualitative research ?   

A: Yes, the following are few common sampling strategies used in qualitative research :  

1. Convenience sampling  

Selects participants who are most easily accessible to researchers due to geographical proximity, availability at a particular time, etc.  

2. Purposive sampling  

Participants are grouped according to predefined criteria based on a specific research question. Sample sizes are often determined based on theoretical saturation (when new data no longer provide additional insights).  

3. Snowball sampling  

Already selected participants use their social networks to refer the researcher to other potential participants.  

4. Quota sampling  

While designing the study, the researchers decide how many people with which characteristics to include as participants. The characteristics help in choosing people most likely to provide insights into the subject.  

qualitative meaning of research

Q: What ethical standards need to be followed with qualitative research ?  

A: The following ethical standards should be considered in qualitative research:  

  • Anonymity : The participants should never be identified in the study and researchers should ensure that no identifying information is mentioned even indirectly.  
  • Confidentiality : To protect participants’ confidentiality, ensure that all related documents, transcripts, notes are stored safely.  
  • Informed consent : Researchers should clearly communicate the objective of the study and how the participants’ responses will be used prior to engaging with the participants.  

Q: How do I address bias in my qualitative research ?  

  A: You could use the following points to ensure an unbiased approach to your qualitative research :  

  • Check your interpretations of the findings with others’ interpretations to identify consistencies.  
  • If possible, you could ask your participants if your interpretations convey their beliefs to a significant extent.  
  • Data triangulation is a way of using multiple data sources to see if all methods consistently support your interpretations.  
  • Contemplate other possible explanations for your findings or interpretations and try ruling them out if possible.  
  • Conduct a peer review of your findings to identify any gaps that may not have been visible to you.  
  • Frame context-appropriate questions to ensure there is no researcher or participant bias.

We hope this article has given you answers to the question “ what is qualitative research ” and given you an in-depth understanding of the various aspects of qualitative research , including the definition, types, and approaches, when to use this method, and advantages and disadvantages, so that the next time you undertake a study you would know which type of research design to adopt.  

References:  

  • McLeod, S. A. Qualitative vs. quantitative research. Simply Psychology [Accessed January 17, 2023]. www.simplypsychology.org/qualitative-quantitative.html    
  • Omniconvert website [Accessed January 18, 2023]. https://www.omniconvert.com/blog/qualitative-research-definition-methodology-limitation-examples/  
  • Busetto L., Wick W., Gumbinger C. How to use and assess qualitative research methods. Neurological Research and Practice [Accessed January 19, 2023] https://neurolrespract.biomedcentral.com/articles/10.1186/s42466-020-00059  
  • QuestionPro website. Qualitative research methods: Types & examples [Accessed January 16, 2023]. https://www.questionpro.com/blog/qualitative-research-methods/  
  • Campuslabs website. How to analyze qualitative data [Accessed January 18, 2023]. https://baselinesupport.campuslabs.com/hc/en-us/articles/204305675-How-to-analyze-qualitative-data  
  • Thematic website. Qualitative data analysis: Step-by-guide [Accessed January 20, 2023]. https://getthematic.com/insights/qualitative-data-analysis/  
  • Lane L. J., Jones D., Penny G. R. Qualitative case study of teachers’ morale in a turnaround school. Research in Higher Education Journal . https://files.eric.ed.gov/fulltext/EJ1233111.pdf  
  • Meetingsnet website. 7 FAQs about qualitative research and CME [Accessed January 21, 2023]. https://www.meetingsnet.com/cme-design/7-faqs-about-qualitative-research-and-cme     
  • Qualitative research methods: A data collector’s field guide. Khoury College of Computer Sciences. Northeastern University. https://course.ccs.neu.edu/is4800sp12/resources/qualmethods.pdf  

Editage All Access is a subscription-based platform that unifies the best AI tools and services designed to speed up, simplify, and streamline every step of a researcher’s journey. The Editage All Access Pack is a one-of-a-kind subscription that unlocks full access to an AI writing assistant, literature recommender, journal finder, scientific illustration tool, and exclusive discounts on professional publication services from Editage.  

Based on 22+ years of experience in academia, Editage All Access empowers researchers to put their best research forward and move closer to success. Explore our top AI Tools pack, AI Tools + Publication Services pack, or Build Your Own Plan. Find everything a researcher needs to succeed, all in one place –  Get All Access now starting at just $14 a month !    

Related Posts

Peer Review Week 2024

Join Us for Peer Review Week 2024

Editage All Access Boosting Productivity for Academics in India

How Editage All Access is Boosting Productivity for Academics in India

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

The PMC website is updating on October 15, 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Neurol Res Pract

Logo of neurrp

How to use and assess qualitative research methods

Loraine busetto.

1 Department of Neurology, Heidelberg University Hospital, Im Neuenheimer Feld 400, 69120 Heidelberg, Germany

Wolfgang Wick

2 Clinical Cooperation Unit Neuro-Oncology, German Cancer Research Center, Heidelberg, Germany

Christoph Gumbinger

Associated data.

Not applicable.

This paper aims to provide an overview of the use and assessment of qualitative research methods in the health sciences. Qualitative research can be defined as the study of the nature of phenomena and is especially appropriate for answering questions of why something is (not) observed, assessing complex multi-component interventions, and focussing on intervention improvement. The most common methods of data collection are document study, (non-) participant observations, semi-structured interviews and focus groups. For data analysis, field-notes and audio-recordings are transcribed into protocols and transcripts, and coded using qualitative data management software. Criteria such as checklists, reflexivity, sampling strategies, piloting, co-coding, member-checking and stakeholder involvement can be used to enhance and assess the quality of the research conducted. Using qualitative in addition to quantitative designs will equip us with better tools to address a greater range of research problems, and to fill in blind spots in current neurological research and practice.

The aim of this paper is to provide an overview of qualitative research methods, including hands-on information on how they can be used, reported and assessed. This article is intended for beginning qualitative researchers in the health sciences as well as experienced quantitative researchers who wish to broaden their understanding of qualitative research.

What is qualitative research?

Qualitative research is defined as “the study of the nature of phenomena”, including “their quality, different manifestations, the context in which they appear or the perspectives from which they can be perceived” , but excluding “their range, frequency and place in an objectively determined chain of cause and effect” [ 1 ]. This formal definition can be complemented with a more pragmatic rule of thumb: qualitative research generally includes data in form of words rather than numbers [ 2 ].

Why conduct qualitative research?

Because some research questions cannot be answered using (only) quantitative methods. For example, one Australian study addressed the issue of why patients from Aboriginal communities often present late or not at all to specialist services offered by tertiary care hospitals. Using qualitative interviews with patients and staff, it found one of the most significant access barriers to be transportation problems, including some towns and communities simply not having a bus service to the hospital [ 3 ]. A quantitative study could have measured the number of patients over time or even looked at possible explanatory factors – but only those previously known or suspected to be of relevance. To discover reasons for observed patterns, especially the invisible or surprising ones, qualitative designs are needed.

While qualitative research is common in other fields, it is still relatively underrepresented in health services research. The latter field is more traditionally rooted in the evidence-based-medicine paradigm, as seen in " research that involves testing the effectiveness of various strategies to achieve changes in clinical practice, preferably applying randomised controlled trial study designs (...) " [ 4 ]. This focus on quantitative research and specifically randomised controlled trials (RCT) is visible in the idea of a hierarchy of research evidence which assumes that some research designs are objectively better than others, and that choosing a "lesser" design is only acceptable when the better ones are not practically or ethically feasible [ 5 , 6 ]. Others, however, argue that an objective hierarchy does not exist, and that, instead, the research design and methods should be chosen to fit the specific research question at hand – "questions before methods" [ 2 , 7 – 9 ]. This means that even when an RCT is possible, some research problems require a different design that is better suited to addressing them. Arguing in JAMA, Berwick uses the example of rapid response teams in hospitals, which he describes as " a complex, multicomponent intervention – essentially a process of social change" susceptible to a range of different context factors including leadership or organisation history. According to him, "[in] such complex terrain, the RCT is an impoverished way to learn. Critics who use it as a truth standard in this context are incorrect" [ 8 ] . Instead of limiting oneself to RCTs, Berwick recommends embracing a wider range of methods , including qualitative ones, which for "these specific applications, (...) are not compromises in learning how to improve; they are superior" [ 8 ].

Research problems that can be approached particularly well using qualitative methods include assessing complex multi-component interventions or systems (of change), addressing questions beyond “what works”, towards “what works for whom when, how and why”, and focussing on intervention improvement rather than accreditation [ 7 , 9 – 12 ]. Using qualitative methods can also help shed light on the “softer” side of medical treatment. For example, while quantitative trials can measure the costs and benefits of neuro-oncological treatment in terms of survival rates or adverse effects, qualitative research can help provide a better understanding of patient or caregiver stress, visibility of illness or out-of-pocket expenses.

How to conduct qualitative research?

Given that qualitative research is characterised by flexibility, openness and responsivity to context, the steps of data collection and analysis are not as separate and consecutive as they tend to be in quantitative research [ 13 , 14 ]. As Fossey puts it : “sampling, data collection, analysis and interpretation are related to each other in a cyclical (iterative) manner, rather than following one after another in a stepwise approach” [ 15 ]. The researcher can make educated decisions with regard to the choice of method, how they are implemented, and to which and how many units they are applied [ 13 ]. As shown in Fig.  1 , this can involve several back-and-forth steps between data collection and analysis where new insights and experiences can lead to adaption and expansion of the original plan. Some insights may also necessitate a revision of the research question and/or the research design as a whole. The process ends when saturation is achieved, i.e. when no relevant new information can be found (see also below: sampling and saturation). For reasons of transparency, it is essential for all decisions as well as the underlying reasoning to be well-documented.

An external file that holds a picture, illustration, etc.
Object name is 42466_2020_59_Fig1_HTML.jpg

Iterative research process

While it is not always explicitly addressed, qualitative methods reflect a different underlying research paradigm than quantitative research (e.g. constructivism or interpretivism as opposed to positivism). The choice of methods can be based on the respective underlying substantive theory or theoretical framework used by the researcher [ 2 ].

Data collection

The methods of qualitative data collection most commonly used in health research are document study, observations, semi-structured interviews and focus groups [ 1 , 14 , 16 , 17 ].

Document study

Document study (also called document analysis) refers to the review by the researcher of written materials [ 14 ]. These can include personal and non-personal documents such as archives, annual reports, guidelines, policy documents, diaries or letters.

Observations

Observations are particularly useful to gain insights into a certain setting and actual behaviour – as opposed to reported behaviour or opinions [ 13 ]. Qualitative observations can be either participant or non-participant in nature. In participant observations, the observer is part of the observed setting, for example a nurse working in an intensive care unit [ 18 ]. In non-participant observations, the observer is “on the outside looking in”, i.e. present in but not part of the situation, trying not to influence the setting by their presence. Observations can be planned (e.g. for 3 h during the day or night shift) or ad hoc (e.g. as soon as a stroke patient arrives at the emergency room). During the observation, the observer takes notes on everything or certain pre-determined parts of what is happening around them, for example focusing on physician-patient interactions or communication between different professional groups. Written notes can be taken during or after the observations, depending on feasibility (which is usually lower during participant observations) and acceptability (e.g. when the observer is perceived to be judging the observed). Afterwards, these field notes are transcribed into observation protocols. If more than one observer was involved, field notes are taken independently, but notes can be consolidated into one protocol after discussions. Advantages of conducting observations include minimising the distance between the researcher and the researched, the potential discovery of topics that the researcher did not realise were relevant and gaining deeper insights into the real-world dimensions of the research problem at hand [ 18 ].

Semi-structured interviews

Hijmans & Kuyper describe qualitative interviews as “an exchange with an informal character, a conversation with a goal” [ 19 ]. Interviews are used to gain insights into a person’s subjective experiences, opinions and motivations – as opposed to facts or behaviours [ 13 ]. Interviews can be distinguished by the degree to which they are structured (i.e. a questionnaire), open (e.g. free conversation or autobiographical interviews) or semi-structured [ 2 , 13 ]. Semi-structured interviews are characterized by open-ended questions and the use of an interview guide (or topic guide/list) in which the broad areas of interest, sometimes including sub-questions, are defined [ 19 ]. The pre-defined topics in the interview guide can be derived from the literature, previous research or a preliminary method of data collection, e.g. document study or observations. The topic list is usually adapted and improved at the start of the data collection process as the interviewer learns more about the field [ 20 ]. Across interviews the focus on the different (blocks of) questions may differ and some questions may be skipped altogether (e.g. if the interviewee is not able or willing to answer the questions or for concerns about the total length of the interview) [ 20 ]. Qualitative interviews are usually not conducted in written format as it impedes on the interactive component of the method [ 20 ]. In comparison to written surveys, qualitative interviews have the advantage of being interactive and allowing for unexpected topics to emerge and to be taken up by the researcher. This can also help overcome a provider or researcher-centred bias often found in written surveys, which by nature, can only measure what is already known or expected to be of relevance to the researcher. Interviews can be audio- or video-taped; but sometimes it is only feasible or acceptable for the interviewer to take written notes [ 14 , 16 , 20 ].

Focus groups

Focus groups are group interviews to explore participants’ expertise and experiences, including explorations of how and why people behave in certain ways [ 1 ]. Focus groups usually consist of 6–8 people and are led by an experienced moderator following a topic guide or “script” [ 21 ]. They can involve an observer who takes note of the non-verbal aspects of the situation, possibly using an observation guide [ 21 ]. Depending on researchers’ and participants’ preferences, the discussions can be audio- or video-taped and transcribed afterwards [ 21 ]. Focus groups are useful for bringing together homogeneous (to a lesser extent heterogeneous) groups of participants with relevant expertise and experience on a given topic on which they can share detailed information [ 21 ]. Focus groups are a relatively easy, fast and inexpensive method to gain access to information on interactions in a given group, i.e. “the sharing and comparing” among participants [ 21 ]. Disadvantages include less control over the process and a lesser extent to which each individual may participate. Moreover, focus group moderators need experience, as do those tasked with the analysis of the resulting data. Focus groups can be less appropriate for discussing sensitive topics that participants might be reluctant to disclose in a group setting [ 13 ]. Moreover, attention must be paid to the emergence of “groupthink” as well as possible power dynamics within the group, e.g. when patients are awed or intimidated by health professionals.

Choosing the “right” method

As explained above, the school of thought underlying qualitative research assumes no objective hierarchy of evidence and methods. This means that each choice of single or combined methods has to be based on the research question that needs to be answered and a critical assessment with regard to whether or to what extent the chosen method can accomplish this – i.e. the “fit” between question and method [ 14 ]. It is necessary for these decisions to be documented when they are being made, and to be critically discussed when reporting methods and results.

Let us assume that our research aim is to examine the (clinical) processes around acute endovascular treatment (EVT), from the patient’s arrival at the emergency room to recanalization, with the aim to identify possible causes for delay and/or other causes for sub-optimal treatment outcome. As a first step, we could conduct a document study of the relevant standard operating procedures (SOPs) for this phase of care – are they up-to-date and in line with current guidelines? Do they contain any mistakes, irregularities or uncertainties that could cause delays or other problems? Regardless of the answers to these questions, the results have to be interpreted based on what they are: a written outline of what care processes in this hospital should look like. If we want to know what they actually look like in practice, we can conduct observations of the processes described in the SOPs. These results can (and should) be analysed in themselves, but also in comparison to the results of the document analysis, especially as regards relevant discrepancies. Do the SOPs outline specific tests for which no equipment can be observed or tasks to be performed by specialized nurses who are not present during the observation? It might also be possible that the written SOP is outdated, but the actual care provided is in line with current best practice. In order to find out why these discrepancies exist, it can be useful to conduct interviews. Are the physicians simply not aware of the SOPs (because their existence is limited to the hospital’s intranet) or do they actively disagree with them or does the infrastructure make it impossible to provide the care as described? Another rationale for adding interviews is that some situations (or all of their possible variations for different patient groups or the day, night or weekend shift) cannot practically or ethically be observed. In this case, it is possible to ask those involved to report on their actions – being aware that this is not the same as the actual observation. A senior physician’s or hospital manager’s description of certain situations might differ from a nurse’s or junior physician’s one, maybe because they intentionally misrepresent facts or maybe because different aspects of the process are visible or important to them. In some cases, it can also be relevant to consider to whom the interviewee is disclosing this information – someone they trust, someone they are otherwise not connected to, or someone they suspect or are aware of being in a potentially “dangerous” power relationship to them. Lastly, a focus group could be conducted with representatives of the relevant professional groups to explore how and why exactly they provide care around EVT. The discussion might reveal discrepancies (between SOPs and actual care or between different physicians) and motivations to the researchers as well as to the focus group members that they might not have been aware of themselves. For the focus group to deliver relevant information, attention has to be paid to its composition and conduct, for example, to make sure that all participants feel safe to disclose sensitive or potentially problematic information or that the discussion is not dominated by (senior) physicians only. The resulting combination of data collection methods is shown in Fig.  2 .

An external file that holds a picture, illustration, etc.
Object name is 42466_2020_59_Fig2_HTML.jpg

Possible combination of data collection methods

Attributions for icons: “Book” by Serhii Smirnov, “Interview” by Adrien Coquet, FR, “Magnifying Glass” by anggun, ID, “Business communication” by Vectors Market; all from the Noun Project

The combination of multiple data source as described for this example can be referred to as “triangulation”, in which multiple measurements are carried out from different angles to achieve a more comprehensive understanding of the phenomenon under study [ 22 , 23 ].

Data analysis

To analyse the data collected through observations, interviews and focus groups these need to be transcribed into protocols and transcripts (see Fig.  3 ). Interviews and focus groups can be transcribed verbatim , with or without annotations for behaviour (e.g. laughing, crying, pausing) and with or without phonetic transcription of dialects and filler words, depending on what is expected or known to be relevant for the analysis. In the next step, the protocols and transcripts are coded , that is, marked (or tagged, labelled) with one or more short descriptors of the content of a sentence or paragraph [ 2 , 15 , 23 ]. Jansen describes coding as “connecting the raw data with “theoretical” terms” [ 20 ]. In a more practical sense, coding makes raw data sortable. This makes it possible to extract and examine all segments describing, say, a tele-neurology consultation from multiple data sources (e.g. SOPs, emergency room observations, staff and patient interview). In a process of synthesis and abstraction, the codes are then grouped, summarised and/or categorised [ 15 , 20 ]. The end product of the coding or analysis process is a descriptive theory of the behavioural pattern under investigation [ 20 ]. The coding process is performed using qualitative data management software, the most common ones being InVivo, MaxQDA and Atlas.ti. It should be noted that these are data management tools which support the analysis performed by the researcher(s) [ 14 ].

An external file that holds a picture, illustration, etc.
Object name is 42466_2020_59_Fig3_HTML.jpg

From data collection to data analysis

Attributions for icons: see Fig. ​ Fig.2, 2 , also “Speech to text” by Trevor Dsouza, “Field Notes” by Mike O’Brien, US, “Voice Record” by ProSymbols, US, “Inspection” by Made, AU, and “Cloud” by Graphic Tigers; all from the Noun Project

How to report qualitative research?

Protocols of qualitative research can be published separately and in advance of the study results. However, the aim is not the same as in RCT protocols, i.e. to pre-define and set in stone the research questions and primary or secondary endpoints. Rather, it is a way to describe the research methods in detail, which might not be possible in the results paper given journals’ word limits. Qualitative research papers are usually longer than their quantitative counterparts to allow for deep understanding and so-called “thick description”. In the methods section, the focus is on transparency of the methods used, including why, how and by whom they were implemented in the specific study setting, so as to enable a discussion of whether and how this may have influenced data collection, analysis and interpretation. The results section usually starts with a paragraph outlining the main findings, followed by more detailed descriptions of, for example, the commonalities, discrepancies or exceptions per category [ 20 ]. Here it is important to support main findings by relevant quotations, which may add information, context, emphasis or real-life examples [ 20 , 23 ]. It is subject to debate in the field whether it is relevant to state the exact number or percentage of respondents supporting a certain statement (e.g. “Five interviewees expressed negative feelings towards XYZ”) [ 21 ].

How to combine qualitative with quantitative research?

Qualitative methods can be combined with other methods in multi- or mixed methods designs, which “[employ] two or more different methods [ …] within the same study or research program rather than confining the research to one single method” [ 24 ]. Reasons for combining methods can be diverse, including triangulation for corroboration of findings, complementarity for illustration and clarification of results, expansion to extend the breadth and range of the study, explanation of (unexpected) results generated with one method with the help of another, or offsetting the weakness of one method with the strength of another [ 1 , 17 , 24 – 26 ]. The resulting designs can be classified according to when, why and how the different quantitative and/or qualitative data strands are combined. The three most common types of mixed method designs are the convergent parallel design , the explanatory sequential design and the exploratory sequential design. The designs with examples are shown in Fig.  4 .

An external file that holds a picture, illustration, etc.
Object name is 42466_2020_59_Fig4_HTML.jpg

Three common mixed methods designs

In the convergent parallel design, a qualitative study is conducted in parallel to and independently of a quantitative study, and the results of both studies are compared and combined at the stage of interpretation of results. Using the above example of EVT provision, this could entail setting up a quantitative EVT registry to measure process times and patient outcomes in parallel to conducting the qualitative research outlined above, and then comparing results. Amongst other things, this would make it possible to assess whether interview respondents’ subjective impressions of patients receiving good care match modified Rankin Scores at follow-up, or whether observed delays in care provision are exceptions or the rule when compared to door-to-needle times as documented in the registry. In the explanatory sequential design, a quantitative study is carried out first, followed by a qualitative study to help explain the results from the quantitative study. This would be an appropriate design if the registry alone had revealed relevant delays in door-to-needle times and the qualitative study would be used to understand where and why these occurred, and how they could be improved. In the exploratory design, the qualitative study is carried out first and its results help informing and building the quantitative study in the next step [ 26 ]. If the qualitative study around EVT provision had shown a high level of dissatisfaction among the staff members involved, a quantitative questionnaire investigating staff satisfaction could be set up in the next step, informed by the qualitative study on which topics dissatisfaction had been expressed. Amongst other things, the questionnaire design would make it possible to widen the reach of the research to more respondents from different (types of) hospitals, regions, countries or settings, and to conduct sub-group analyses for different professional groups.

How to assess qualitative research?

A variety of assessment criteria and lists have been developed for qualitative research, ranging in their focus and comprehensiveness [ 14 , 17 , 27 ]. However, none of these has been elevated to the “gold standard” in the field. In the following, we therefore focus on a set of commonly used assessment criteria that, from a practical standpoint, a researcher can look for when assessing a qualitative research report or paper.

Assessors should check the authors’ use of and adherence to the relevant reporting checklists (e.g. Standards for Reporting Qualitative Research (SRQR)) to make sure all items that are relevant for this type of research are addressed [ 23 , 28 ]. Discussions of quantitative measures in addition to or instead of these qualitative measures can be a sign of lower quality of the research (paper). Providing and adhering to a checklist for qualitative research contributes to an important quality criterion for qualitative research, namely transparency [ 15 , 17 , 23 ].

Reflexivity

While methodological transparency and complete reporting is relevant for all types of research, some additional criteria must be taken into account for qualitative research. This includes what is called reflexivity, i.e. sensitivity to the relationship between the researcher and the researched, including how contact was established and maintained, or the background and experience of the researcher(s) involved in data collection and analysis. Depending on the research question and population to be researched this can be limited to professional experience, but it may also include gender, age or ethnicity [ 17 , 27 ]. These details are relevant because in qualitative research, as opposed to quantitative research, the researcher as a person cannot be isolated from the research process [ 23 ]. It may influence the conversation when an interviewed patient speaks to an interviewer who is a physician, or when an interviewee is asked to discuss a gynaecological procedure with a male interviewer, and therefore the reader must be made aware of these details [ 19 ].

Sampling and saturation

The aim of qualitative sampling is for all variants of the objects of observation that are deemed relevant for the study to be present in the sample “ to see the issue and its meanings from as many angles as possible” [ 1 , 16 , 19 , 20 , 27 ] , and to ensure “information-richness [ 15 ]. An iterative sampling approach is advised, in which data collection (e.g. five interviews) is followed by data analysis, followed by more data collection to find variants that are lacking in the current sample. This process continues until no new (relevant) information can be found and further sampling becomes redundant – which is called saturation [ 1 , 15 ] . In other words: qualitative data collection finds its end point not a priori , but when the research team determines that saturation has been reached [ 29 , 30 ].

This is also the reason why most qualitative studies use deliberate instead of random sampling strategies. This is generally referred to as “ purposive sampling” , in which researchers pre-define which types of participants or cases they need to include so as to cover all variations that are expected to be of relevance, based on the literature, previous experience or theory (i.e. theoretical sampling) [ 14 , 20 ]. Other types of purposive sampling include (but are not limited to) maximum variation sampling, critical case sampling or extreme or deviant case sampling [ 2 ]. In the above EVT example, a purposive sample could include all relevant professional groups and/or all relevant stakeholders (patients, relatives) and/or all relevant times of observation (day, night and weekend shift).

Assessors of qualitative research should check whether the considerations underlying the sampling strategy were sound and whether or how researchers tried to adapt and improve their strategies in stepwise or cyclical approaches between data collection and analysis to achieve saturation [ 14 ].

Good qualitative research is iterative in nature, i.e. it goes back and forth between data collection and analysis, revising and improving the approach where necessary. One example of this are pilot interviews, where different aspects of the interview (especially the interview guide, but also, for example, the site of the interview or whether the interview can be audio-recorded) are tested with a small number of respondents, evaluated and revised [ 19 ]. In doing so, the interviewer learns which wording or types of questions work best, or which is the best length of an interview with patients who have trouble concentrating for an extended time. Of course, the same reasoning applies to observations or focus groups which can also be piloted.

Ideally, coding should be performed by at least two researchers, especially at the beginning of the coding process when a common approach must be defined, including the establishment of a useful coding list (or tree), and when a common meaning of individual codes must be established [ 23 ]. An initial sub-set or all transcripts can be coded independently by the coders and then compared and consolidated after regular discussions in the research team. This is to make sure that codes are applied consistently to the research data.

Member checking

Member checking, also called respondent validation , refers to the practice of checking back with study respondents to see if the research is in line with their views [ 14 , 27 ]. This can happen after data collection or analysis or when first results are available [ 23 ]. For example, interviewees can be provided with (summaries of) their transcripts and asked whether they believe this to be a complete representation of their views or whether they would like to clarify or elaborate on their responses [ 17 ]. Respondents’ feedback on these issues then becomes part of the data collection and analysis [ 27 ].

Stakeholder involvement

In those niches where qualitative approaches have been able to evolve and grow, a new trend has seen the inclusion of patients and their representatives not only as study participants (i.e. “members”, see above) but as consultants to and active participants in the broader research process [ 31 – 33 ]. The underlying assumption is that patients and other stakeholders hold unique perspectives and experiences that add value beyond their own single story, making the research more relevant and beneficial to researchers, study participants and (future) patients alike [ 34 , 35 ]. Using the example of patients on or nearing dialysis, a recent scoping review found that 80% of clinical research did not address the top 10 research priorities identified by patients and caregivers [ 32 , 36 ]. In this sense, the involvement of the relevant stakeholders, especially patients and relatives, is increasingly being seen as a quality indicator in and of itself.

How not to assess qualitative research

The above overview does not include certain items that are routine in assessments of quantitative research. What follows is a non-exhaustive, non-representative, experience-based list of the quantitative criteria often applied to the assessment of qualitative research, as well as an explanation of the limited usefulness of these endeavours.

Protocol adherence

Given the openness and flexibility of qualitative research, it should not be assessed by how well it adheres to pre-determined and fixed strategies – in other words: its rigidity. Instead, the assessor should look for signs of adaptation and refinement based on lessons learned from earlier steps in the research process.

Sample size

For the reasons explained above, qualitative research does not require specific sample sizes, nor does it require that the sample size be determined a priori [ 1 , 14 , 27 , 37 – 39 ]. Sample size can only be a useful quality indicator when related to the research purpose, the chosen methodology and the composition of the sample, i.e. who was included and why.

Randomisation

While some authors argue that randomisation can be used in qualitative research, this is not commonly the case, as neither its feasibility nor its necessity or usefulness has been convincingly established for qualitative research [ 13 , 27 ]. Relevant disadvantages include the negative impact of a too large sample size as well as the possibility (or probability) of selecting “ quiet, uncooperative or inarticulate individuals ” [ 17 ]. Qualitative studies do not use control groups, either.

Interrater reliability, variability and other “objectivity checks”

The concept of “interrater reliability” is sometimes used in qualitative research to assess to which extent the coding approach overlaps between the two co-coders. However, it is not clear what this measure tells us about the quality of the analysis [ 23 ]. This means that these scores can be included in qualitative research reports, preferably with some additional information on what the score means for the analysis, but it is not a requirement. Relatedly, it is not relevant for the quality or “objectivity” of qualitative research to separate those who recruited the study participants and collected and analysed the data. Experiences even show that it might be better to have the same person or team perform all of these tasks [ 20 ]. First, when researchers introduce themselves during recruitment this can enhance trust when the interview takes place days or weeks later with the same researcher. Second, when the audio-recording is transcribed for analysis, the researcher conducting the interviews will usually remember the interviewee and the specific interview situation during data analysis. This might be helpful in providing additional context information for interpretation of data, e.g. on whether something might have been meant as a joke [ 18 ].

Not being quantitative research

Being qualitative research instead of quantitative research should not be used as an assessment criterion if it is used irrespectively of the research problem at hand. Similarly, qualitative research should not be required to be combined with quantitative research per se – unless mixed methods research is judged as inherently better than single-method research. In this case, the same criterion should be applied for quantitative studies without a qualitative component.

The main take-away points of this paper are summarised in Table ​ Table1. 1 . We aimed to show that, if conducted well, qualitative research can answer specific research questions that cannot to be adequately answered using (only) quantitative designs. Seeing qualitative and quantitative methods as equal will help us become more aware and critical of the “fit” between the research problem and our chosen methods: I can conduct an RCT to determine the reasons for transportation delays of acute stroke patients – but should I? It also provides us with a greater range of tools to tackle a greater range of research problems more appropriately and successfully, filling in the blind spots on one half of the methodological spectrum to better address the whole complexity of neurological research and practice.

Take-away-points

• Assessing complex multi-component interventions or systems (of change)

• What works for whom when, how and why?

• Focussing on intervention improvement

• Document study

• Observations (participant or non-participant)

• Interviews (especially semi-structured)

• Focus groups

• Transcription of audio-recordings and field notes into transcripts and protocols

• Coding of protocols

• Using qualitative data management software

• Combinations of quantitative and/or qualitative methods, e.g.:

• : quali and quanti in parallel

• : quanti followed by quali

• : quali followed by quanti

• Checklists

• Reflexivity

• Sampling strategies

• Piloting

• Co-coding

• Member checking

• Stakeholder involvement

• Protocol adherence

• Sample size

• Randomization

• Interrater reliability, variability and other “objectivity checks”

• Not being quantitative research

Acknowledgements

Abbreviations.

EVTEndovascular treatment
RCTRandomised Controlled Trial
SOPStandard Operating Procedure
SRQRStandards for Reporting Qualitative Research

Authors’ contributions

LB drafted the manuscript; WW and CG revised the manuscript; all authors approved the final versions.

no external funding.

Availability of data and materials

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

The authors declare no competing interests.

Publisher’s Note

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

Logo for JCU Open eBooks

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

4.2 Definitions and Characteristics of Qualitative Research

Qualitative research aims to uncover the meaning and understanding of phenomena that cannot be broken down into measurable elements. It is based on naturalistic, interpretative and humanistic notions. 5 This research method seeks to discover, explore, identify or describe subjective human experiences using non-statistical methods and develops themes from the study participants’ stories. 5 Figure 4.1 depicts major features/ characteristics of qualitative research. It utilises exploratory open-ended questions and observations to search for patterns of meaning in collected data (e.g. observation, verbal/written narrative data, photographs, etc.) and uses inductive thinking (from specific observations to more general rules) to interpret meaning. 6 Participants’ voice is evident through quotations and description of the work. 6 The context/ setting of the study and the researcher’s reflexivity (i.e. “reflection on and awareness of their bias”, the effect of the researcher’s experience on the data and interpretations) are very important and described as part of data collection. 6 Analysis of collected data is complex, often involves inductive data analysis (exploration, contrasts, specific to general) and requires multiple coding and development of themes from participant stories. 6

flow chart of characteristics of qualitative research

Reflexivity- avoiding bias/Role of the qualitative researcher

Qualitative researchers generally begin their work with the recognition that their position (or worldview) has a significant impact on the overall research process. 7 Researcher worldview shapes the way the research is conducted, i.e., how the questions are formulated, methods are chosen, data are collected and analysed, and results are reported. Therefore, it is essential for qualitative researchers to acknowledge, articulate, reflect on and clarify their own underlying biases and assumptions before embarking on any research project. 7 Reflexivity helps to ensure that the researcher’s own experiences, values, and beliefs do not unintentionally bias the data collection, analysis, and interpretation. 7 It is the gold standard for establishing trustworthiness and has been established as one of the ways qualitative researchers should ensure rigour and quality in their work. 8 The following questions in Table 4.1 may help you begin the reflective process. 9

Table 4.1: Questions to aid the reflection process

What piques my interest in this subject? You need to consider what motivates your excitement, energy, and interest in investigating this topic to answer this question
What exactly do I believe the solution is? Asking this question allows you to detect any biases by honestly reflecting on what you anticipate finding. The assumptions can be grouped/classified to allow the participants’ opinions to be heard.
What exactly am I getting out of this? In many circumstances, the “pressure to publish” reduces research to nothing more than a job necessity. What effect does this have on your interest in the subject and its results? To what extent are you willing to go to find information?
What do my colleagues think of this project—and me? You will not work in a vacuum as a researcher; you will be part of a social and interpersonal world. These outside factors will impact your perceptions of yourself and your job.

Recognising this impact and its possible implications on human behaviour will allow for more self-reflection during the study process.

Philosophical underpinnings to qualitative research

Qualitative research uses an inductive approach and stems from interpretivism or constructivism and assumes that realities are multiple, socially constructed, and holistic. 10 According to this philosophical viewpoint, humans build reality through their interactions with the world around them. 10 As a result, qualitative research aims to comprehend how individuals make sense of their experiences and build meaning in their lives. 10 Because reality is complex/nuanced and context-bound, participants constantly construct it depending on their understanding. Thus, the interactions between the researcher and the participants are considered necessary to offer a rich description of the concept and provide an in-depth understanding of the phenomenon under investigation. 11

An Introduction to Research Methods for Undergraduate Health Profession Students Copyright © 2023 by Faith Alele and Bunmi Malau-Aduli is licensed under a Creative Commons Attribution-NonCommercial 4.0 International License , except where otherwise noted.

An Overview of Qualitative Research Methods

Direct Observation, Interviews, Participation, Immersion, Focus Groups

  • Research, Samples, and Statistics
  • Key Concepts
  • Major Sociologists
  • News & Issues
  • Recommended Reading
  • Archaeology

Qualitative research is a type of social science research that collects and works with non-numerical data and that seeks to interpret meaning from these data that help understand social life through the study of targeted populations or places.

People often frame it in opposition to quantitative research , which uses numerical data to identify large-scale trends and employs statistical operations to determine causal and correlative relationships between variables.

Within sociology, qualitative research is typically focused on the micro-level of social interaction that composes everyday life, whereas quantitative research typically focuses on macro-level trends and phenomena.

Key Takeaways

Methods of qualitative research include:

  • observation and immersion
  • open-ended surveys
  • focus groups
  • content analysis of visual and textual materials
  • oral history

Qualitative research has a long history in sociology and has been used within it for as long as the field has existed.

This type of research has long appealed to social scientists because it allows the researchers to investigate the meanings people attribute to their behavior, actions, and interactions with others.

While quantitative research is useful for identifying relationships between variables, like, for example, the connection between poverty and racial hate, it is qualitative research that can illuminate why this connection exists by going directly to the source—the people themselves.

Qualitative research is designed to reveal the meaning that informs the action or outcomes that are typically measured by quantitative research. So qualitative researchers investigate meanings, interpretations, symbols, and the processes and relations of social life.

What this type of research produces is descriptive data that the researcher must then interpret using rigorous and systematic methods of transcribing, coding, and analysis of trends and themes.

Because its focus is everyday life and people's experiences, qualitative research lends itself well to creating new theories using the inductive method , which can then be tested with further research.

Qualitative researchers use their own eyes, ears, and intelligence to collect in-depth perceptions and descriptions of targeted populations, places, and events.

Their findings are collected through a variety of methods, and often a researcher will use at least two or several of the following while conducting a qualitative study:

  • Direct observation : With direct observation, a researcher studies people as they go about their daily lives without participating or interfering. This type of research is often unknown to those under study, and as such, must be conducted in public settings where people do not have a reasonable expectation of privacy. For example, a researcher might observe the ways in which strangers interact in public as they gather to watch a street performer.
  • Open-ended surveys : While many surveys are designed to generate quantitative data, many are also designed with open-ended questions that allow for the generation and analysis of qualitative data. For example, a survey might be used to investigate not just which political candidates voters chose, but why they chose them, in their own words.
  • Focus group : In a focus group, a researcher engages a small group of participants in a conversation designed to generate data relevant to the research question. Focus groups can contain anywhere from 5 to 15 participants. Social scientists often use them in studies that examine an event or trend that occurs within a specific community. They are common in market research, too.
  • In-depth interviews : Researchers conduct in-depth interviews by speaking with participants in a one-on-one setting. Sometimes a researcher approaches the interview with a predetermined list of questions or topics for discussion but allows the conversation to evolve based on how the participant responds. Other times, the researcher has identified certain topics of interest but does not have a formal guide for the conversation, but allows the participant to guide it.
  • Oral history : The oral history method is used to create a historical account of an event, group, or community, and typically involves a series of in-depth interviews conducted with one or multiple participants over an extended period.
  • Participant observation : This method is similar to observation, however with this one, the researcher also participates in the action or events to not only observe others but to gain the first-hand experience in the setting.
  • Ethnographic observation : Ethnographic observation is the most intensive and in-depth observational method. Originating in anthropology, with this method, a researcher fully immerses themselves into the research setting and lives among the participants as one of them for anywhere from months to years. By doing this, the researcher attempts to experience day-to-day existence from the viewpoints of those studied to develop in-depth and long-term accounts of the community, events, or trends under observation.
  • Content analysis : This method is used by sociologists to analyze social life by interpreting words and images from documents, film, art, music, and other cultural products and media. The researchers look at how the words and images are used, and the context in which they are used to draw inferences about the underlying culture. Content analysis of digital material, especially that generated by social media users, has become a popular technique within the social sciences.

While much of the data generated by qualitative research is coded and analyzed using just the researcher's eyes and brain, the use of computer software to do these processes is increasingly popular within the social sciences.

Such software analysis works well when the data is too large for humans to handle, though the lack of a human interpreter is a common criticism of the use of computer software.

Pros and Cons

Qualitative research has both benefits and drawbacks.

On the plus side, it creates an in-depth understanding of the attitudes, behaviors, interactions, events, and social processes that comprise everyday life. In doing so, it helps social scientists understand how everyday life is influenced by society-wide things like social structure , social order , and all kinds of social forces.

This set of methods also has the benefit of being flexible and easily adaptable to changes in the research environment and can be conducted with minimal cost in many cases.

Among the downsides of qualitative research is that its scope is fairly limited so its findings are not always widely able to be generalized.

Researchers also have to use caution with these methods to ensure that they do not influence the data in ways that significantly change it and that they do not bring undue personal bias to their interpretation of the findings.

Fortunately, qualitative researchers receive rigorous training designed to eliminate or reduce these types of research bias.

  • How to Conduct a Sociology Research Interview
  • What Is Participant Observation Research?
  • Immersion Definition: Cultural, Language, and Virtual
  • Definition and Overview of Grounded Theory
  • The Differences Between Indexes and Scales
  • Pros and Cons of Secondary Data Analysis
  • Social Surveys: Questionnaires, Interviews, and Telephone Polls
  • The Different Types of Sampling Designs in Sociology
  • Principal Components and Factor Analysis
  • Sociology Explains Why Some People Cheat on Their Spouses
  • Deductive Versus Inductive Reasoning
  • How to Construct an Index for Research
  • Data Sources For Sociological Research
  • A Review of Software Tools for Quantitative Data Analysis
  • Constructing a Deductive Theory
  • Scales Used in Social Science Research

What is qualitative research?

The most fundamental characteristic of qualitative research is its express commitment to viewing events, action, norms, values, etc. from the perspective of the people who experience them in everyday life. (Bryman, 2004: p. 61)

The term “qualitative research” refers to an umbrella concept that encompasses many different forms of inquiry and methodological practices. It engages a variety of theoretical lenses, strategies, and techniques. Different from quantitative research, which is based on probability and measurement ( quantity ), qualitative research is based on the quality of the data generated to explain a phenomenon (e.g., why older adults would resist using some kinds of mobility devices) (Gardner, 2014). Traditionally, in the health sciences, qualitative research has been defined in opposition to quantitative research. A stereotypical view of qualitative research is that it is defined by its data generation methods, such as interviews and observations. These techniques are in fact shared with quantitative research; for example in psychological studies, observation is a commonly used strategy for quantification of behaviours (Green & Thorogood, 2004).  We caution against these simplistic generalizations. As illustrated in this chapter, all qualitative health methodologies and methods are centred around the notion of knowledge production grounded in the quality (the explanatory potential) of the information generated about a phenomenon.

Qualitative research is best defined by its aims: it asks different questions and has a different focus than quantitative research. It is concerned with questions of how, why, and what (Green & Thorogood, 2004).  Qualitative research is rooted in the social sciences and is concerned with people and their social realities (Bryman, 2004), with how the social world is understood, experienced, interpreted, and constituted; with individual and collective meanings, interpretations, practices/behaviours, and social processes. Its perspective is emic; it focuses on the subjectivity of human experiences (de la Cuesta, 2015). In the health sciences, qualitative research is the ideal approach for studying the meanings people give to their experiences and how they make sense of their social worlds (e.g., patients’ perceptions of self-care education or reasons for adherence, or not, to prescribed medication). Health care and health promotion are largely shaped by people’s perceptions, social norms, and organizational standards and practices; all these issues are social in nature and hence can be studied qualitatively.

Qualitative research is also based on a naturalistic approach to data generation. This means that people, situations, and events are studied where they happen, in their “natural settings,” and thus all qualitative data are contextual, connected to the people, places, times, events, and the everyday social interactions – or “social and cultural contexts” –  in which the data are generated. Context is also essential for understanding social behaviours and for making sense of or analyzing data produced. Context includes considerations such as who, when, where, why, class, race/ethnicity/gender, age, and circumstances (Holstein & Gubrium, 2004; Korstjens & Moser, 2017).

Qualitative research has also been described in terms of its broad purposes or goals: exploratory, where researchers investigate phenomena about which little is known; explanatory, where relationships, events, behaviours, or beliefs related to a group are explained; descriptive, where experiences or events are documented; and emancipatory, where the goal is to create opportunities for people to engage in social action (Agee, 2009). While these objectives can be helpful in situating a study, they are artificial distinctions because qualitative researchers often combine more than one goal in their study design. For example, answering a question such as “what are the processes that shape the ability of patients with diabetes to follow a prescribed diet?” requires both description and explanation.

Additionally, qualitative research traditions vary according to the uses researchers in distinct disciplines make of them. Within health sciences, for example, there is a particular way to think about “types” of qualitative research. Eakin (2016) refers to the dominance of post-positivist qualitative research (PPQR) in the health sciences, where qualitative data is viewed through a positivist lens: for instance, data are “real” and speak for themselves; findings “emerge” from the data independently of the researcher, who assumes a veneer of neutrality; and “findings” are reported mainly in implicitly quantitative terms (“some,” “most”). In PPQR, qualitative research is conceived of as purely a “method or technique, a ‘toolbox’ of procedures divorced from their philosophical undercarriage” (Eakin, 2016: p.111). Eakin concludes that this type of qualitative research has “limited value either as positivist or interpretive enterprise [because] it cannot satisfy the criteria for adequate positivist design (objective standardized procedure, statistical generalizability) or for adequate interpretive design (researcher as instrument, conceptual generalizability)” (p. 111).

As we are interested in doing the most we can when we use qualitative research to improve health care delivery and to challenge the ways we think about health issues broadly (e.g., social discourses, policies, programs), we engage with an interpretive, rather than a post-positivist, form of qualitative research. This is also called “interpretive research” (Schwartz-Shea & Yanow, 2012). Interpretive qualitative research is rooted in the assumption that meaning is discerned by the researcher. Through language use, human interaction, and meaning-making, the researcher and participant create the conditions for an in-depth understanding of a phenomenon (e.g., the stigma associated with TB treatment).

This type of interpretive qualitative research strives for what Eakin (2016)  calls “value-added” analysis. It is an approach that refuses a mere cataloguing of pre-conceived or common-sense ideas, maximizes the “creative presence” of the researcher, and deploys theoretical abstraction as a key methodological strategy for reconceptualizing phenomena and creating generalizable knowledge, through the process of theorization (Eakin, 2016). Here the researcher goes beyond mere description of an experience or a phenomenon to question, for instance, commonly held notions and assumptions, or the everyday experiences that are taken for granted. The researcher does not take data as given but works hard to interpret it, considers the story behind the story, questions common-sense and received understandings, and asks questions about the nature of the phenomenon under study (Eakin, 2016). (see also Ward, Hoare & Gott, 2015).

Creative Commons License

Share This Book

  • Increase Font Size

Qualitative vs Quantitative Research Methods & Data Analysis

Saul McLeod, PhD

Editor-in-Chief for Simply Psychology

BSc (Hons) Psychology, MRes, PhD, University of Manchester

Saul McLeod, PhD., is a qualified psychology teacher with over 18 years of experience in further and higher education. He has been published in peer-reviewed journals, including the Journal of Clinical Psychology.

Learn about our Editorial Process

Olivia Guy-Evans, MSc

Associate Editor for Simply Psychology

BSc (Hons) Psychology, MSc Psychology of Education

Olivia Guy-Evans is a writer and associate editor for Simply Psychology. She has previously worked in healthcare and educational sectors.

The main difference between quantitative and qualitative research is the type of data they collect and analyze.

Quantitative data is information about quantities, and therefore numbers, and qualitative data is descriptive, and regards phenomenon which can be observed but not measured, such as language.
  • Quantitative research collects numerical data and analyzes it using statistical methods. The aim is to produce objective, empirical data that can be measured and expressed numerically. Quantitative research is often used to test hypotheses, identify patterns, and make predictions.
  • Qualitative research gathers non-numerical data (words, images, sounds) to explore subjective experiences and attitudes, often via observation and interviews. It aims to produce detailed descriptions and uncover new insights about the studied phenomenon.

On This Page:

What Is Qualitative Research?

Qualitative research is the process of collecting, analyzing, and interpreting non-numerical data, such as language. Qualitative research can be used to understand how an individual subjectively perceives and gives meaning to their social reality.

Qualitative data is non-numerical data, such as text, video, photographs, or audio recordings. This type of data can be collected using diary accounts or in-depth interviews and analyzed using grounded theory or thematic analysis.

Qualitative research is multimethod in focus, involving an interpretive, naturalistic approach to its subject matter. This means that qualitative researchers study things in their natural settings, attempting to make sense of, or interpret, phenomena in terms of the meanings people bring to them. Denzin and Lincoln (1994, p. 2)

Interest in qualitative data came about as the result of the dissatisfaction of some psychologists (e.g., Carl Rogers) with the scientific study of psychologists such as behaviorists (e.g., Skinner ).

Since psychologists study people, the traditional approach to science is not seen as an appropriate way of carrying out research since it fails to capture the totality of human experience and the essence of being human.  Exploring participants’ experiences is known as a phenomenological approach (re: Humanism ).

Qualitative research is primarily concerned with meaning, subjectivity, and lived experience. The goal is to understand the quality and texture of people’s experiences, how they make sense of them, and the implications for their lives.

Qualitative research aims to understand the social reality of individuals, groups, and cultures as nearly as possible as participants feel or live it. Thus, people and groups are studied in their natural setting.

Some examples of qualitative research questions are provided, such as what an experience feels like, how people talk about something, how they make sense of an experience, and how events unfold for people.

Research following a qualitative approach is exploratory and seeks to explain ‘how’ and ‘why’ a particular phenomenon, or behavior, operates as it does in a particular context. It can be used to generate hypotheses and theories from the data.

Qualitative Methods

There are different types of qualitative research methods, including diary accounts, in-depth interviews , documents, focus groups , case study research , and ethnography .

The results of qualitative methods provide a deep understanding of how people perceive their social realities and in consequence, how they act within the social world.

The researcher has several methods for collecting empirical materials, ranging from the interview to direct observation, to the analysis of artifacts, documents, and cultural records, to the use of visual materials or personal experience. Denzin and Lincoln (1994, p. 14)

Here are some examples of qualitative data:

Interview transcripts : Verbatim records of what participants said during an interview or focus group. They allow researchers to identify common themes and patterns, and draw conclusions based on the data. Interview transcripts can also be useful in providing direct quotes and examples to support research findings.

Observations : The researcher typically takes detailed notes on what they observe, including any contextual information, nonverbal cues, or other relevant details. The resulting observational data can be analyzed to gain insights into social phenomena, such as human behavior, social interactions, and cultural practices.

Unstructured interviews : generate qualitative data through the use of open questions.  This allows the respondent to talk in some depth, choosing their own words.  This helps the researcher develop a real sense of a person’s understanding of a situation.

Diaries or journals : Written accounts of personal experiences or reflections.

Notice that qualitative data could be much more than just words or text. Photographs, videos, sound recordings, and so on, can be considered qualitative data. Visual data can be used to understand behaviors, environments, and social interactions.

Qualitative Data Analysis

Qualitative research is endlessly creative and interpretive. The researcher does not just leave the field with mountains of empirical data and then easily write up his or her findings.

Qualitative interpretations are constructed, and various techniques can be used to make sense of the data, such as content analysis, grounded theory (Glaser & Strauss, 1967), thematic analysis (Braun & Clarke, 2006), or discourse analysis .

For example, thematic analysis is a qualitative approach that involves identifying implicit or explicit ideas within the data. Themes will often emerge once the data has been coded .

RESEARCH THEMATICANALYSISMETHOD

Key Features

  • Events can be understood adequately only if they are seen in context. Therefore, a qualitative researcher immerses her/himself in the field, in natural surroundings. The contexts of inquiry are not contrived; they are natural. Nothing is predefined or taken for granted.
  • Qualitative researchers want those who are studied to speak for themselves, to provide their perspectives in words and other actions. Therefore, qualitative research is an interactive process in which the persons studied teach the researcher about their lives.
  • The qualitative researcher is an integral part of the data; without the active participation of the researcher, no data exists.
  • The study’s design evolves during the research and can be adjusted or changed as it progresses. For the qualitative researcher, there is no single reality. It is subjective and exists only in reference to the observer.
  • The theory is data-driven and emerges as part of the research process, evolving from the data as they are collected.

Limitations of Qualitative Research

  • Because of the time and costs involved, qualitative designs do not generally draw samples from large-scale data sets.
  • The problem of adequate validity or reliability is a major criticism. Because of the subjective nature of qualitative data and its origin in single contexts, it is difficult to apply conventional standards of reliability and validity. For example, because of the central role played by the researcher in the generation of data, it is not possible to replicate qualitative studies.
  • Also, contexts, situations, events, conditions, and interactions cannot be replicated to any extent, nor can generalizations be made to a wider context than the one studied with confidence.
  • The time required for data collection, analysis, and interpretation is lengthy. Analysis of qualitative data is difficult, and expert knowledge of an area is necessary to interpret qualitative data. Great care must be taken when doing so, for example, looking for mental illness symptoms.

Advantages of Qualitative Research

  • Because of close researcher involvement, the researcher gains an insider’s view of the field. This allows the researcher to find issues that are often missed (such as subtleties and complexities) by the scientific, more positivistic inquiries.
  • Qualitative descriptions can be important in suggesting possible relationships, causes, effects, and dynamic processes.
  • Qualitative analysis allows for ambiguities/contradictions in the data, which reflect social reality (Denscombe, 2010).
  • Qualitative research uses a descriptive, narrative style; this research might be of particular benefit to the practitioner as she or he could turn to qualitative reports to examine forms of knowledge that might otherwise be unavailable, thereby gaining new insight.

What Is Quantitative Research?

Quantitative research involves the process of objectively collecting and analyzing numerical data to describe, predict, or control variables of interest.

The goals of quantitative research are to test causal relationships between variables , make predictions, and generalize results to wider populations.

Quantitative researchers aim to establish general laws of behavior and phenomenon across different settings/contexts. Research is used to test a theory and ultimately support or reject it.

Quantitative Methods

Experiments typically yield quantitative data, as they are concerned with measuring things.  However, other research methods, such as controlled observations and questionnaires , can produce both quantitative information.

For example, a rating scale or closed questions on a questionnaire would generate quantitative data as these produce either numerical data or data that can be put into categories (e.g., “yes,” “no” answers).

Experimental methods limit how research participants react to and express appropriate social behavior.

Findings are, therefore, likely to be context-bound and simply a reflection of the assumptions that the researcher brings to the investigation.

There are numerous examples of quantitative data in psychological research, including mental health. Here are a few examples:

Another example is the Experience in Close Relationships Scale (ECR), a self-report questionnaire widely used to assess adult attachment styles .

The ECR provides quantitative data that can be used to assess attachment styles and predict relationship outcomes.

Neuroimaging data : Neuroimaging techniques, such as MRI and fMRI, provide quantitative data on brain structure and function.

This data can be analyzed to identify brain regions involved in specific mental processes or disorders.

For example, the Beck Depression Inventory (BDI) is a clinician-administered questionnaire widely used to assess the severity of depressive symptoms in individuals.

The BDI consists of 21 questions, each scored on a scale of 0 to 3, with higher scores indicating more severe depressive symptoms. 

Quantitative Data Analysis

Statistics help us turn quantitative data into useful information to help with decision-making. We can use statistics to summarize our data, describing patterns, relationships, and connections. Statistics can be descriptive or inferential.

Descriptive statistics help us to summarize our data. In contrast, inferential statistics are used to identify statistically significant differences between groups of data (such as intervention and control groups in a randomized control study).

  • Quantitative researchers try to control extraneous variables by conducting their studies in the lab.
  • The research aims for objectivity (i.e., without bias) and is separated from the data.
  • The design of the study is determined before it begins.
  • For the quantitative researcher, the reality is objective, exists separately from the researcher, and can be seen by anyone.
  • Research is used to test a theory and ultimately support or reject it.

Limitations of Quantitative Research

  • Context: Quantitative experiments do not take place in natural settings. In addition, they do not allow participants to explain their choices or the meaning of the questions they may have for those participants (Carr, 1994).
  • Researcher expertise: Poor knowledge of the application of statistical analysis may negatively affect analysis and subsequent interpretation (Black, 1999).
  • Variability of data quantity: Large sample sizes are needed for more accurate analysis. Small-scale quantitative studies may be less reliable because of the low quantity of data (Denscombe, 2010). This also affects the ability to generalize study findings to wider populations.
  • Confirmation bias: The researcher might miss observing phenomena because of focus on theory or hypothesis testing rather than on the theory of hypothesis generation.

Advantages of Quantitative Research

  • Scientific objectivity: Quantitative data can be interpreted with statistical analysis, and since statistics are based on the principles of mathematics, the quantitative approach is viewed as scientifically objective and rational (Carr, 1994; Denscombe, 2010).
  • Useful for testing and validating already constructed theories.
  • Rapid analysis: Sophisticated software removes much of the need for prolonged data analysis, especially with large volumes of data involved (Antonius, 2003).
  • Replication: Quantitative data is based on measured values and can be checked by others because numerical data is less open to ambiguities of interpretation.
  • Hypotheses can also be tested because of statistical analysis (Antonius, 2003).

Antonius, R. (2003). Interpreting quantitative data with SPSS . Sage.

Black, T. R. (1999). Doing quantitative research in the social sciences: An integrated approach to research design, measurement and statistics . Sage.

Braun, V. & Clarke, V. (2006). Using thematic analysis in psychology . Qualitative Research in Psychology , 3, 77–101.

Carr, L. T. (1994). The strengths and weaknesses of quantitative and qualitative research : what method for nursing? Journal of advanced nursing, 20(4) , 716-721.

Denscombe, M. (2010). The Good Research Guide: for small-scale social research. McGraw Hill.

Denzin, N., & Lincoln. Y. (1994). Handbook of Qualitative Research. Thousand Oaks, CA, US: Sage Publications Inc.

Glaser, B. G., Strauss, A. L., & Strutzel, E. (1968). The discovery of grounded theory; strategies for qualitative research. Nursing research, 17(4) , 364.

Minichiello, V. (1990). In-Depth Interviewing: Researching People. Longman Cheshire.

Punch, K. (1998). Introduction to Social Research: Quantitative and Qualitative Approaches. London: Sage

Further Information

  • Mixed methods research
  • Designing qualitative research
  • Methods of data collection and analysis
  • Introduction to quantitative and qualitative research
  • Checklists for improving rigour in qualitative research: a case of the tail wagging the dog?
  • Qualitative research in health care: Analysing qualitative data
  • Qualitative data analysis: the framework approach
  • Using the framework method for the analysis of
  • Qualitative data in multi-disciplinary health research
  • Content Analysis
  • Grounded Theory
  • Thematic Analysis

Print Friendly, PDF & Email

Forage

Qualitative Research Definition

Qualitative research methods and examples, advantages and disadvantages of qualitative approaches, qualitative vs. quantitative research, showing qualitative research skills on resumes, what is qualitative research methods and examples.

McKayla Girardin

  • Share on Twitter Share on Twitter
  • Share on Facebook Share on Facebook
  • Share on LinkedIn Share on LinkedIn

What Is Qualitative Research? Examples and methods

Forage puts students first. Our blog articles are written independently by our editorial team. They have not been paid for or sponsored by our partners. See our full  editorial guidelines .

Table of Contents

Qualitative research seeks to understand people’s experiences and perspectives by studying social organizations and human behavior. Data in qualitative studies focuses on people’s beliefs and emotional responses. Qualitative data is especially helpful when a company wants to know how customers feel about a product or service, such as in user experience (UX) design or marketing . 

Researchers use qualitative approaches to “determine answers to research questions on human behavior and the cultural values that drive our thinking and behavior,” says Margaret J. King, director at The Center for Cultural Studies & Analysis in Philadelphia.

Data in qualitative research typically can’t be assessed mathematically — the data is not sets of numbers or quantifiable information. Rather, it’s collections of images, words, notes on behaviors, descriptions of emotions, and historical context. Data is collected through observations, interviews, surveys, focus groups, and secondary research. 

However, a qualitative study needs a “clear research question at its base,” notes King, and the research needs to be “observed, categorized, compared, and evaluated (along a scale or by a typology chart) by reference to a baseline in order to determine an outcome with value as new and reliable information.”

What's the difference: Data science vs. software engineering?

Quantium Data Analytics

Explore the power of data and its ability to power breakthrough possibilities for individuals, organisations and societies with this free job simulation from Quantium.

Avg. Time: 4 to 5 hours

Skills you’ll build: Data validation, data visualisation, data wrangling, programming, data analysis, commercial thinking, statistical testing, presentation skills

Who Uses Qualitative Research?

Researchers in social sciences and humanities often use qualitative research methods, especially in specific areas of study like anthropology, history, education, and sociology. 

Qualitative methods are also applicable in business, technology , and marketing spaces. For example, product managers use qualitative research to understand how target audiences respond to their products. They may use focus groups to gain insights from potential customers on product prototypes and improvements or surveys from existing customers to understand what changes users want to see. 

Other careers that may involve qualitative research include: 

  • Marketing analyst
  • UX and UI analyst
  • Market researcher
  • Statistician
  • Business analyst
  • Data analyst
  • Research assistant
  • Claims investigator

Working at Electronic Arts

Electronic Arts Product Management

Learn how product managers leverage qualitative and other types of research to build and improve products in this free job simulation from EA.

Avg. Time: 1 to 2 hours

Skills you’ll build: Critical thinking, problem solving, performance metrics, written communication, project planning

Good research begins with a question, and this question informs the approach used by qualitative researchers. 

Grounded Theory

Grounded theory is an inductive approach to theory development. In many forms of research, you begin with a hypothesis and then test it to see if you’re correct. In grounded theory, though, you go in without any assumptions and rely on the data you collect to form theories. You start with an open question about a phenomenon you are studying and collect and analyze data until you can form a fully-fledged theory from the information. 

Example: A company wants to improve its brand and marketing strategies. The company performs a grounded theory approach to solving this problem by conducting interviews and surveys with past, current, and prospective customers. The information gathered from these methods helps the company understand what type of branding and marketing their customer-base likes and dislikes, allowing the team to inductively craft a new brand and marketing strategy from the data. 

Action Research

Action research is one part study and one part problem-solving . Through action research, analysts investigate a problem or weakness and develop practical solutions. The process of action research is cyclical —- researchers assess solutions for efficiency and effectiveness, and create further solutions to correct any issues found. 

Example: A manager notices her employees struggle to cooperate on group projects. She carefully reviews how team members interact with each other and asks them all to respond to a survey about communication. Through the survey and study, she finds that guidelines for group projects are unclear. After changing the guidelines, she reviews her team again to see if there are any changes to their behavior.  

>>MORE: Explore how action research helps consultants serve clients with Accenture’s Client Research and Problem Identification job simulation .

Phenomenological Research

Phenomenological research investigates a phenomenon in depth, looking at people’s experiences and understanding of the situation. This sort of study is primarily descriptive and seeks to broaden understanding around a specific incident and the people involved. Researchers in phenomenological studies must be careful to set aside any biases or assumptions because the information used should be entirely from the subjects themselves. 

Example : A researcher wants to better understand the lived experience of college students with jobs. The purpose of this research is to gain insights into the pressures of college students who balance studying and working at the same time. The researcher conducts a series of interviews with several college students, learning about their past and current situations. Through the first few interviews, the researcher builds a relationship with the students. Later discussions are more targeted, with questions prompting the students to discuss their emotions surrounding both work and school and the difficulties and benefits arising from their situation. The researcher then analyzes these interviews, and identifies shared themes to contextualize the experiences of the students.

qualitative meaning of research

GE Aerospace Human Resources

Learn the research and conflict resolution skills necessary for a career in human resources in this free job simulation from GE.

Avg. Time: 3 to 4 hours

Skills you’ll build: Feedback giving, communication skills, empowering with insights, basics of lean, process mapping, continuous improvement tools, dataset handling in Excel

Ethnography

Ethnography is an immersive study of a particular culture or community. Through ethnographic research, analysts aim to learn about a group’s conventions, social dynamics, and cultural norms. Some researchers use active observation methods, finding ways to integrate themselves into the culture as much as possible. Others use passive observation, watching closely from the outside but not fully immersing themselves. 

Example: A company hires an external researcher to learn what their company’s culture is actually like. The researcher studies the social dynamics of the employees and may even look at how these employees interact with clients and with each other outside of the office. The goal is to deliver a comprehensive report of the company’s culture and the social dynamics of its employees.

Case Studies

A case study is a type of in-depth analysis of a situation. Case studies can focus on an organization, belief system, event, person, or action. The goal of a case study is to understand the phenomenon and put it in a real-world context. Case studies are also commonly used in marketing and sales to highlight the benefits of a company’s products or services. 

Example: A business performs a case study of its competitors’ strategies. This case study aims to show why the company should adopt a specific business strategy. The study looks at each competitor’s business structure, marketing campaigns, product offerings, and historical growth trends. Then, using this data on other businesses, the researcher can theorize how that strategy would benefit their company.

>>MORE: Learn how companies use case study interviews to assess candidates’ research and problem-solving skills. 

Qualitative research methods are great for generating new ideas. The exploratory nature of qualitative research means uncovering unexpected information, which often leads to new theories and further research topics. Additionally, qualitative findings feel meaningful. These studies focus on people, emotions, and societies and may feel closer to their communities than quantitative research that relies on more mathematical and logical data. 

However, qualitative research can be unreliable at times. It’s difficult to replicate qualitative studies since people’s opinions and emotions can change quickly. For example, a focus group has a lot of variables that can affect the outcome, and that same group, asked the same questions a year later, may have entirely different responses. The data collection can also be difficult and time-consuming with qualitative research. Ultimately, interviewing people, reviewing surveys, and understanding and explaining human emotions can be incredibly complex.

Find your career fit

See what career path is right for you with our free career quiz!

While qualitative research deals with data that isn’t easily manipulated by mathematics, quantitative research almost exclusively involves numbers and numerical data. Quantitative studies aim to find concrete details, like units of time, percentages, or statistics. 

Besides the types of data used, a core difference between quantitative and qualitative research is the idea of control and replication. 

“Qualitative is less subject to control (as in lab studies) and, therefore, less statistically measurable than quantitative approaches,” says King.

One person’s interview about a specific topic can have completely different responses than every other person’s interview since there are so many variables in qualitative research. On the other hand, quantitative studies can often be replicated. For instance, when testing the effects of a new medication, quantifiable data, like blood test results, can be repeated. Qualitative data, though, like how people feel about the medication, may differ from person to person and from moment to moment.

J.P.Morgan bank corporation headquarters glass building concept. JP Morgan banking company symbol on front facade 3d illustration.

JPMorgan Quantitative Research

Discover how bankers use quantitative methods to analyze businesses and industry trends with this free job simulation.

Avg. Time: 6 to 7 hours

Skills you’ll build: Programming, data analysis, Python, critical thinking, statistics, dynamic programming

You can show your experience with qualitative research on your resume in your skills or work experience sections and your cover letter . 

  • In your skills section , you can list types of qualitative research you are skilled at, like conducting interviews, performing grounded theory research, or crafting case studies. 
  • In your work or internship experience descriptions , you can highlight specific examples, like talking about a time you used action research to solve a complex issue at your last job. 
  • In your cover letter , you can discuss in-depth qualitative research projects you’ve completed. For instance, say you spent a summer conducting ethnographic research or a whole semester running focus groups to get feedback on a product. You can talk about these experiences in your cover letter and note how these skills make you a great fit for the job. 

Grow your skills and explore your career options with Forage’s free job simulations .

Image credit: Canva

McKayla Girardin

Related Posts

6 negotiation skills to level up your work life, how to build conflict resolution skills: case studies and examples, what is github uses and getting started, upskill with forage.

Bloomberg logo through magnifying glass

Top companies are hiring now!

Logo for VCU Pressbooks

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

29 Conceptualization in qualitative research

Chapter outline

  • 15.1 Alternative paradigms: Interpretivism, critical paradigm, and pragmatism

15.2 Multiparadigmatic research: An example

15.3 idiographic causal relationships, 15.4 qualitative research questions.

Now let’s change things up! In the previous chapters, we explored steps to create and carry out a quantitative research study. Quantitative studies are great when we want to summarize or test relationships between ideas using numbers and the power of statistics. However, qualitative research offers us a different and equally important tool. Sometimes the aim of research projects is to explore meaning and lived experience. Instead of trying to arrive at generalizable conclusions for all people, some research projects establish a deep, authentic description of a specific time, place, and group of people.

Qualitative research relies on the power of human expression through words, pictures, movies, performance and other artifacts that represent these things. All of these tell stories about the human experience and we want to learn from them and have them be represented in our research. Generally speaking, qualitative research is about the gathering up of these stories, breaking them into pieces so we can examine the ideas that make them up, and putting them back together in a way that allows us to tell a common or shared story that responds to our research question. To do that, we need to discuss the assumptions underlying social science.

A penguin on an ice float. The top of the float is labeled method, next down is methodology, theory, and philosophical foundations.

17.1 Alternative paradigms: Interpretivism, critical, and pragmatism

Learning objectives.

Students will be able to…

  • Distinguish between the assumptions of positivism, interpretivism, critical, and pragmatist research paradigms.
  • Use paradigm to describe how scientific thought changes over time.

In Chapter 10, we reviewed the assumptions that underly post-positivism (abbreviated hereafter as positivism for brevity). Quantitative methods are most often the choice for positivist research questions because they conform to these assumptions. Qualitative methods  can conform to these assumptions; however, they are limited in their generalizability.

Kivunja & Kuyini (2017) [1] describe the essential features of positivism as:

  • A belief that theory is universal and law-like generalizations can be made across contexts
  • The assumption that context is not important
  • The belief that truth or knowledge is ‘out there to be discovered’ by research
  • The belief that cause and effect are distinguishable and analytically separable
  • The belief that results of inquiry can be quantified
  • The belief that theory can be used to predict and to control outcomes
  • The belief that research should follow the scientific method of investigation
  • Rests on formulation and testing of hypotheses
  • Employs empirical or analytical approaches
  • Pursues an objective search for facts
  • Believes in ability to observe knowledge
  • The researcher’s ultimate aim is to establish a comprehensive universal theory, to account for human and social behavior
  • Application of the scientific method

Because positivism is the dominant social science research paradigm, it can be easy to ignore or be confused by research that does not use these assumptions. We covered in Chapter 10 the table reprinted below when discussing the assumptions underlying positivistic social science.

As you consider your research project, keep these philosophical assumptions in mind. They are useful shortcuts to understanding the deeper ideas and assumptions behind the construction of knowledge. The purpose of exploring these philosophical assumptions isn’t to find out which is true and which is false. Instead, the goal is to identify the assumptions that fit with how you think about your research question. Choosing a paradigm helps you make those assumptions explicit.

Table 7.1 Philosophical assumptions in social science research
Ontology: assumptions about what is real
Epistemology: assumptions about how we come to know what is real

Assumptions about the researcher

Assumptions about human action

Assumptions about the social world
Assumptions about the purpose of research

Before we explore alternative paradigms, it’s important for us to review what paradigms are.

How do scientific ideas change over time?

Much like your ideas develop over time as you learn more, so does the body of scientific knowledge. Kuhn’s (1962) [2] The Structure of Scientific Revolutions is one of the most influential works on the philosophy of science, and is credited with introducing the idea of competing paradigms (or “disciplinary matrices”) in research. Kuhn investigated the way that scientific practices evolve over time, arguing that we don’t have a simple progression from “less knowledge” to “more knowledge” because the way that we approach inquiry is changing over time. This can happen gradually, but the process results in moments of change where our understanding of a phenomenon changes more radically (such as in the transition from Newtonian to Einsteinian physics; or from Lamarckian to Darwinian theories of evolution). For a social work practice example, Fleuridas & Krafcik (2019) [3] trace the development of the “four forces” of psychotherapy , from psychodynamics to behaviorism to humanism as well as the competition among emerging perspectives to establish itself as the fourth force to guide psychotherapeutic practice. But how did the problems in one paradigm inspire new paradigms? Kuhn presents us with a way of understanding the history of scientific development across all topics and disciplines.

As you can see in this video from Matthew J. Brown (CC-BY), there are four stages in the cycle of science in Kuhn’s approach. Firstly, a pre-paradigmatic state where competing approaches share no consensus. Secondly, the “normal” state where there is wide acceptance of a particular set of methods and assumptions. Thirdly, a state of crisis where anomalies that cannot be solved within the existing paradigm emerge and competing theories to address them follow. Fourthly, a revolutionary phase where some new paradigmatic approach becomes dominant and supplants the old. Shnieder (2009) [4] suggests that the Kuhnian phases are characterized by different kinds of scientific activity.

Newer approaches often build upon rather than replace older ones, but they also overlap and can exist within a state of competition. Scientists working within a particular paradigm often share methods, assumptions and values. In addition to supporting specific methods, research paradigms also influence things like the ambition and nature of research, the researcher-participant relationship and how the role of the researcher is understood.

Paradigm vs. theory

The terms ‘ paradigm ‘ and ‘ theory ‘ are often used interchangeably in social science. There is not a consensus among social scientists as to whether these are identical or distinct concepts. With that said, in this text, we will make a clear distinction between the two ideas because thinking about each concept separately is more useful for our purposes.

We define paradigm a set of common philosophical (ontological, epistemological, and axiological) assumptions that inform research. The four paradigms we describe in this section refer to patterns in how groups of researchers resolve philosophical questions. Some assumptions naturally make sense together, and paradigms grow out of researchers with shared assumptions about what is important and how to study it. Paradigms are like “analytic lenses” and a provide framework on top of which we can build theoretical and empirical knowledge (Kuhn, 1962). [5] Consider this video of an interview with world-famous physicist Richard Feynman in which he explains why “when you explain a ‘why,’ you have to be in some framework that you allow something to be true. Otherwise, you are perpetually asking why.” In order to answer basic physics question like “what is happening when two magnets attract?” or a social work research question like “what is the impact of this therapeutic intervention on depression,” you must understand the assumptions you are making about social science and the social world. Paradigmatic assumptions about objective and subjective truth support methodological choices like whether to conduct interviews or send out surveys, for example.

While paradigms are broad philosophical assumptions, theory is more specific, and refers to a set of concepts and relationships scientists use to explain the social world. Theories are more concrete, while paradigms are more abstract. Look back to Figure 7.1 at the beginning of this chapter. Theory helps you identify the concepts and relationships that align with your paradigmatic understanding of the problem. Moreover, theory informs how you will measure the concepts in your research question and the design of your project.

For both theories and paradigms, Kuhn’s observation of scientific paradigms, crises, and revolutions is instructive for understanding the history of science. Researchers inherit institutions, norms, and ideas that are marked by the battlegrounds of theoretical and paradigmatic debates that stretch back hundreds of years. We have necessarily simplified this history into four paradigms: positivism, interpretivism, critical, and pragmatism. Our framework and explanation are inspired by the framework of Guba and Lincoln (1990) [6] and Burrell and Morgan (1979). [7] while also incorporating pragmatism as a way of resolving paradigmatic questions. Most of social work research and theory can be classified as belonging to one of these four paradigms, though this classification system represents only one of many useful approaches to analyzing social science research paradigms.

Building on our discussion in section 7.1 on objective vs. subjective epistemologies and ontologies, we will start with the difference between positivism and interpretivism. Afterward, we will link our discussion of axiology in section 7.2 with the critical paradigm. Finally, we will situate pragmatism as a way to resolve paradigmatic questions strategically. The difference between positivism and interpretivism is a good place to start, since the critical paradigm and pragmatism build on their philosophical insights.

It’s important to think of paradigms less as distinct categories and more as a spectrum along which projects might fall. For example, some projects may be somewhat positivist, somewhat interpretivist, and a little critical. No project fits perfectly into one paradigm. Additionally, there is no paradigm that is more correct than the other. Each paradigm uses assumptions that are logically consistent, and when combined, are a useful approach to understanding the social world using science. The purpose of this section is to acquaint you with what research projects in each paradigm look like and how they are grounded in philosophical assumptions about social science.

You should read this section to situate yourself in terms of what paradigm feels most “at home” to both you as a person and to your project. You may find, as I have, that your research projects are more conventional and less radical than what feels most like home to you, personally. In a research project, however, students should start with their working question rather than their heart. Use the paradigm that fits with your question the best, rather than which paradigm you think fits you the best.

qualitative meaning of research

Interpretivism: Researcher as “empathizer”

Positivism is focused on generalizable truth. Interpretivism , by contrast, develops from the idea that we want to understand the truths of individuals, how they interpret and experience the world, their thought processes, and the social structures that emerge from sharing those interpretations through language and behavior. The process of interpretation (or social construction) is guided by the empathy of the researcher to understand the meaning behind what other people say.

Historically, interpretivism grew out of a specific critique of positivism: that knowledge in the human and social sciences cannot conform to the model of natural science because there are features of human experience that cannot objectively be “known”. The tools we use to understand objects that have no self-awareness may not be well-attuned to subjective experiences like emotions, understandings, values, feelings, socio-cultural factors, historical influences, and other meaningful aspects of social life. Instead of finding a single generalizable “truth,” the interpretivist researcher aims to generate understanding and often adopts a relativist position.

While positivists seek “the truth,” the social constructionist framework argues that “truth” varies. Truth differs based on who you ask, and people change what they believe is true based on social interactions. These subjective truths also exist within social and historical contexts, and our understanding of truth varies across communities and time periods. This is because we, according to this paradigm, create reality ourselves through our social interactions and our interpretations of those interactions. Key to the interpretivist perspective is the idea that social context and interaction frame our realities.

Researchers operating within this framework take keen interest in how people come to socially agree, or disagree, about what is real and true. Consider how people, depending on their social and geographical context, ascribe different meanings to certain hand gestures. When a person raises their middle finger, those of us in Western cultures will probably think that this person isn’t very happy (not to mention the person at whom the middle finger is being directed!). In other societies around the world, a thumbs-up gesture, rather than a middle finger, signifies discontent (Wong, 2007). [8] The fact that these hand gestures have different meanings across cultures aptly demonstrates that those meanings are socially and collectively constructed. What, then, is the “truth” of the middle finger or thumbs up? As we’ve seen in this section, the truth depends on the intention of the person making the gesture, the interpretation of the person receiving it, and the social context in which the action occurred.

Qualitative methods are preferred as ways to investigate these phenomena. Data collected might be unstructured (or “messy”) and correspondingly a range of techniques for approaching data collection have been developed. Interpretivism acknowledges that it is impossible to remove cultural and individual influence from research, often instead making a virtue of the positionality of the researcher and the socio-cultural context of a study.

One common objection positivists levy against interpretivists is that interpretivism tends to emphasize the subjective over the objective. If the starting point for an investigation is that we can’t fully and objectively know the world, how can we do research into this without everything being a matter of opinion? For the positivist, this risk for confirmation bias as well as invalid and unreliable measures makes interpretivist research unscientific. Clearly, we disagree with this assessment, and you should, too. Positivism and interpretivism have different ontologies and epistemologies with contrasting notions of rigor and validity (for more information on assumptions about measurement, see Chapter 11 for quantitative validity and reliability and Chapter 20 for qualitative rigor). Nevertheless, both paradigms apply the values and concepts of the scientific method through systematic investigation of the social world, even if their assumptions lead them to do so in different ways. Interpretivist research often embraces a relativist epistemology, bringing together different perspectives in search of a trustworthy and authentic understanding or narrative.

Kivunja & Kuyini (2017) [9] describe the essential features of interpretivism as:

  • The belief that truths are multiple and socially constructed
  • The acceptance that there is inevitable interaction between the researcher and his or her research participants
  • The acceptance that context is vital for knowledge and knowing
  • The belief that knowledge can be value laden and the researcher’s values need to be made explicit
  • The need to understand specific cases and contexts rather deriving universal laws that apply to everyone, everywhere.
  • The belief that causes and effects are mutually interdependent, and that causality may be circular or contradictory
  • The belief that contextual factors need to be taken into consideration in any systematic pursuit of understanding

One important clarification: it’s important to think of the interpretivist perspective as not just about individual interpretations but the social life of interpretations. While individuals may construct their own realities, groups—from a small one such as a married couple to large ones such as nations—often agree on notions of what is true and what “is” and what “is not.” In other words, the meanings that we construct have power beyond the individuals who create them. Therefore, the ways that people and communities act based on such meanings is of as much interest to interpretivists as how they were created in the first place. Theories like social constructionism, phenomenology, and symbolic interactionism are often used in concert with interpretivism.

Is interpretivism right for your project?

An interpretivist orientation to research is appropriate when your working question asks about subjective truths. The cause-and-effect relationships that interpretivist studies produce are specific to the time and place in which the study happened, rather than a generalizable objective truth. More pragmatically, if you picture yourself having a conversation with participants like an interview or focus group, then interpretivism is likely going to be a major influence for your study.

Positivists critique the interpretivist paradigm as non-scientific. They view the interpretivist focus on subjectivity and values as sources of bias. Positivists and interpretivists differ on the degree to which social phenomena are like natural phenomena. Positivists believe that the assumptions of the social sciences and natural sciences are the same, while interpretivists strongly believe that social sciences differ from the natural sciences because their subjects are social creatures.

Similarly, the critical paradigm finds fault with the interpretivist focus on the status quo rather than social change. Although interpretivists often proceed from a feminist or other standpoint theory, the focus is less on liberation than on understanding the present from multiple perspectives. Other critical theorists may object to the consensus orientation of interpretivist research. By searching for commonalities between people’s stories, they may erase the uniqueness of each individual’s story. For example, while interpretivists may arrive at a consensus definition of what the experience of “coming out” is like for people who identify as lesbian, gay, bisexual, transgender, or queer, it cannot represent the diversity of each person’s unique “coming out” experience and what it meant to them. For example, see Rosario and colleagues’ (2009) [10] critique the literature on lesbians “coming out” because previous studies did not addressing how appearing, behaving, or identifying as a butch or femme impacted the experience of “coming out” for lesbians.

  • From your literature search, identify an empirical article that uses qualitative methods to answer a research question similar to your working question or about your research topic.
  • Review the assumptions of the interpretivist research paradigm.
  • Discuss in a few sentences how the author’s conclusions are based on some of these paradigmatic assumptions. How might a researcher operating from a different paradigm (like positivism or the critical paradigm) critique the conclusions of this study?

qualitative meaning of research

Critical paradigm: Researcher as “activist”

As we’ve discussed a bit in the preceding sections, the critical paradigm focuses on power, inequality, and social change. Although some rather diverse perspectives are included here, the critical paradigm, in general, includes ideas developed by early social theorists, such as Max Horkheimer (Calhoun et al., 2007), [11] and later works developed by feminist scholars, such as Nancy Fraser (1989). [12] Unlike the positivist paradigm, the critical paradigm assumes that social science can never be truly objective or value-free. Furthermore, this paradigm operates from the perspective that scientific investigation should be conducted with the express goal of social change. Researchers in the critical paradigm foreground axiology, positionality and values . In contrast with the detached, “objective” observations associated with the positivist researcher, critical approaches make explicit the intention for research to act as a transformative or emancipatory force within and beyond the study.

Researchers in the critical paradigm might start with the knowledge that systems are biased against certain groups, such as women or ethnic minorities, building upon previous theory and empirical data. Moreover, their research projects are designed not only to collect data, but to impact the participants as well as the systems being studied. The critical paradigm applies its study of power and inequality to change those power imbalances as part of the research process itself. If this sounds familiar to you, you may remember hearing similar ideas when discussing social conflict theory in your human behavior in the social environment (HBSE) class. [13] Because of this focus on social change, the critical paradigm is a natural home for social work research. However, we fall far short of adopting this approach widely in our profession’s research efforts.

Is the critical paradigm right for your project?

Every social work research project impacts social justice in some way. What distinguishes critical research is how it integrates an analysis of power into the research process itself. Critical research is appropriate for projects that are activist in orientation. For example, critical research projects should have working questions that explicitly seek to raise the consciousness of an oppressed group or collaborate equitably with community members and clients to addresses issues of concern. Because of their transformative potential, critical research projects can be incredibly rewarding to complete. However, partnerships take a long time to develop and social change can evolve slowly on an issue, making critical research projects a more challenging fit for student research projects which must be completed under a tight deadline with few resources.

Positivists critique the critical paradigm on multiple fronts. First and foremost, the focus on oppression and values as part of the research process is seen as likely to bias the research process, most problematically, towards confirmation bias. If you start out with the assumption that oppression exists and must be dealt with, then you are likely to find that regardless of whether it is truly there or not. Similarly, positivists may fault critical researchers for focusing on how the world should be, rather than how it truly is . In this, they may focus too much on theoretical and abstract inquiry and less on traditional experimentation and empirical inquiry. Finally, the goal of social transformation is seen as inherently unscientific, as science is not a political practice.

Interpretivists often find common cause with critical researchers. Feminist studies, for example, may explore the perspectives of women while centering gender-based oppression as part of the research process. In interpretivist research, the focus is less on radical change as part of the research process and more on small, incremental changes based on the results and conclusions drawn from the research project. Additionally, some critical researchers’ focus on individuality of experience is in stark contrast to the consensus-orientation of interpretivists. Interpretivists seek to understand people’s true selves. Some critical theorists argue that people have multiple selves or no self at all.

  • From your literature search, identify an article relevant to your working question or broad research topic that uses a critical perspective. You should look for articles where the authors are clear that they are applying a critical approach to research like feminism, anti-racism, Marxism and critical theory, decolonization, anti-oppressive practice, or other social justice-focused theoretical perspectives. To target your search further, include keywords in your queries to research methods commonly used in the critical paradigm like participatory action research and community-based participatory research. If you have trouble identifying an article for this exercise, consult your professor for some help. These articles may be more challenging to find, but reviewing one is necessary to get a feel for what research in this paradigm is like.
  • Review the assumptions of the critical research paradigm.
  • Discuss in a few sentences how the author’s conclusions are based on some of these paradigmatic assumptions. How might a researcher operating from different assumptions (like values-neutrality or researcher as neutral and unbiased) critique the conclusions of this study?

qualitative meaning of research

Pragmatism: Researcher as “strategist”

“Essentially, all models are wrong but some are useful.” (Box, 1976) [14]

Pragmatism is a research paradigm that suspends questions of philosophical ‘truth’ and focuses more on how different philosophies, theories, and methods can be used strategically to provide a multidimensional view of a topic. Researchers employing pragmatism will mix elements of positivist, interpretivist, and critical research depending on the purpose of a particular project and the practical constraints faced by the researcher and their research context. We favor this approach for student projects because it avoids getting bogged down in choosing the “right” paradigm and instead focuses on the assumptions that help you answer your question, given the limitations of your research context. Student research projects are completed quickly and moving in the direction of pragmatism can be a route to successfully completing a project. Your project is a representation of what you think is feasible, ethical, and important enough for you to study.

The crucial consideration for the pragmatist is whether the outcomes of research have any real-world application, rather than whether they are “true.” The methods, theories, and philosophies chosen by pragmatic researchers are guided by their working question. There are no distinctively pragmatic research methods since this approach is about making judicious use whichever methods fit best with the problem under investigation. Pragmatic approaches may be less likely to prioritize ontological, epistemological or axiological consistency when combining different research methods. Instead, the emphasis is on solving a pressing problem and adapting to the limitations and opportunities in the researchers’ context.

Adopt a multi-paradigmatic perspective

Believe it or not, there is a long literature of acrimonious conflict between scientists from positivist, interpretivist, and critical camps (see Heineman-Pieper et al., 2002 [15] for a longer discussion). Pragmatism is an old idea, but it is appealing precisely because it attempts to resolve the problem of multiple incompatible philosophical assumptions in social science. To a pragmatist, there is no “correct” paradigm. All paradigms rely on assumptions about the social world that are the subject of philosophical debate. Each paradigm is an incomplete understanding of the world, and it requires a scientific community using all of them to gain a comprehensive view of the social world. This multi-paradigmatic perspective is a unique gift of social work research, as our emphasis on empathy and social change makes us more critical of positivism, the dominant paradigm in social science.

We offered the metaphors of expert, empathizer, activist, and strategist for each paradigm. It’s important not to take these labels too seriously. For example, some may view that scientists should be experts or that activists are biased and unscientific. Nevertheless, we hope that these metaphors give you a sense of what it feels like to conduct research within each paradigm.

One of the unique aspects of paradigmatic thinking is that often where you think you are most at home may actually be the opposite of where your research project is. For example, in my graduate and doctoral education, I thought I was a critical researcher. In fact, I thought I was a radical researcher focused on social change and transformation. Yet, often times when I sit down to conceptualize and start a research project, I find myself squarely in the positivist paradigm, thinking through neat cause-and-effect relationships that can be mathematically measured. There is nothing wrong with that! Your task for your research project is to find the paradigm that best matches your research question. Think through what you really want to study and how you think about the topic, then use assumptions of that paradigm to guide your inquiry.

Another important lesson is that no research project fits perfectly in one paradigm or another. Instead, there is a spectrum along which studies are, to varying degrees, interpretivist, positivist, and critical. For example, all social work research is a bit activist in that our research projects are designed to inform action for change on behalf of clients and systems. However, some projects will focus on the conclusions and implications of projects informing social change (i.e., positivist and interpretivist projects) while others will partner with community members and design research projects collaboratively in a way that leads to social change (i.e. critical projects). In section 7.5, we will describe a pragmatic approach to research design guided by your paradigmatic and theoretical framework.

Key Takeaways

  • Social work research falls, to some degree, in each of the four paradigms: positivism, interpretivism, critical, and pragmatist.
  • Adopting a pragmatic, multi-paradigmatic approach to research makes sense for student researchers, as it directs students to use the philosophical assumptions and methodological approaches that best match their research question and research context.
  • Research in all paradigms is necessary to come to a comprehensive understanding of a topic, and social workers must be able to understand and apply knowledge from each research paradigm.
  • Describe which paradigm best fits your perspective on the world and which best fits with your project.
  • Identify any similarities and differences in your personal assumptions and the assumption your research project relies upon. For example, are you a more critical and radical thinker but have chosen a more “expert” role for yourself in your research project?

Learners will be able to…

  • Apply the assumptions of each paradigm to your project
  • Summarize what aspects of your project stem from positivist, interpretivist, or critical assumptions

In the previous sections, we reviewed the major paradigms and theories in social work research. In this section, we will provide an example of how to apply theory and paradigm in research. This process is depicted in Figure 7.2 below with some quick summary questions for each stage. Some questions in the figure below have example answers like designs (i.e., experimental, survey) and data analysis approaches (i.e., discourse analysis). These examples are arbitrary. There are a lot of options that are not listed. So, don’t feel like you have to memorize them or use them in your study.

qualitative meaning of research

This diagram (taken from an archived Open University (UK) course entitled E89 ​- Educational Inquiry ) ​ shows one way to visualize the research design process. While research is far from linear, in general, this is how research projects progress sequentially. Researchers begin with a working question, and through engaging with the literature, develop and refine those questions into research questions (a process we will finalize in Chapter 9 ). But in order to get to the part where you gather your sample, measure your participants, and analyze your data, you need to start with paradigm. Based on your work in section 7.3, you should have a sense of which paradigm or paradigms are best suited to answering your question. The approach taken will often reflect the nature of the research question; the kind of data it is possible to collect; and work previously done in the area under consideration. When evaluating paradigm and theory, it is important to look at what other authors have done previously and the framework used by studies that are similar to the one you are thinking of conducting.

Once you situate your project in a research paradigm, it becomes possible to start making concrete choices about methods. Depending on the project, this will involve choices about things like:

  • What is my final research question?
  • What are the key variables and concepts under investigation, and how will I measure them?
  • How do I find a representative sample of people who experience the topic I’m studying?
  • What design is most appropriate for my research question?
  • How will I collect and analyze data?
  • How do I determine whether my results describe real patterns in the world or are the result of bias or error?

The data collection phase can begin once these decisions are made. It can be very tempting to start collecting data as soon as possible in the research process as this gives a sense of progress. However, it is usually worth getting things exactly right before collecting data as an error found in your approach further down the line can be harder to correct or recalibrate around.

Designing a study using paradigm and theory: An example

Paradigm and theory have the potential to turn some people off since there is a lot of abstract terminology and thinking about real-world social work practice contexts. In this section, I’ll use an example from my own research, and I hope it will illustrate a few things. First, it will show that paradigms are really just philosophical statements about things you already understand and think about normally. It will also show that no project neatly sits in one paradigm and that a social work researcher should use whichever paradigm or combination of paradigms suit their question the best. Finally, I hope it is one example of how to be a pragmatist and strategically use the strengths of different theories and paradigms to answering a research question. We will pick up the discussion of mixed methods in the next chapter.

Thinking as an expert: Positivism

In my undergraduate research methods class, I used an open textbook much like this one and wanted to study whether it improved student learning. You can read a copy of the article we wrote on based on our study . We’ll learn more about the specifics of experiments and evaluation research in Chapter 13 , but you know enough to understand what evaluating an intervention might look like. My first thought was to conduct an experiment, which placed me firmly within the positivist or “expert” paradigm.

Experiments focus on isolating the relationship between cause and effect. For my study, this meant studying an open textbook (the cause, or intervention) and final grades (the effect, or outcome). Notice that my position as “expert” lets me assume many things in this process. First, it assumes that I can distill the many dimensions of student learning into one number—the final grade. Second, as the “expert,” I’ve determined what the intervention is: indeed, I created the book I was studying, and applied a theory from experts in the field that explains how and why it should impact student learning.

Theory is part of applying all paradigms, but I’ll discuss its impact within positivism first. Theories grounded in positivism help explain why one thing causes another. More specifically, these theories isolate a causal relationship between two (or more) concepts while holding constant the effects of other variables that might confound the relationship between the key variables. That is why experimental design is so common in positivist research. The researcher isolates the environment from anything that might impact or bias the cause and effect relationship they want to investigate.

But in order for one thing to lead to change in something else, there must be some logical, rational reason why it would do so. In open education, there are a few hypotheses (though no full-fledged theories) on why students might perform better using open textbooks. The most common is the access hypothesis , which states that students who cannot afford expensive textbooks or wouldn’t buy them anyway can access open textbooks because they are free, which will improve their grades. It’s important to note that I held this theory prior to starting the experiment, as in positivist research you spell out your hypotheses in advance and design an experiment to support or refute that hypothesis.

Notice that the hypothesis here applies not only to the people in my experiment, but to any student in higher education. Positivism seeks generalizable truth, or what is true for everyone. The results of my study should provide evidence that  anyone  who uses an open textbook would achieve similar outcomes. Of course, there were a number of limitations as it was difficult to tightly control the study. I could not randomly assign students or prevent them from sharing resources with one another, for example. So, while this study had many positivist elements, it was far from a perfect positivist study because I was forced to adapt to the pragmatic limitations of my research context (e.g., I cannot randomly assign students to classes) that made it difficult to establish an objective, generalizable truth.

Thinking like an empathizer: Interpretivism

One of the things that did not sit right with me about the study was the reliance on final grades to signify everything that was going on with students. I added another quantitative measure that measured research knowledge, but this was still too simplistic. I wanted to understand how students used the book and what they thought about it. I could create survey questions that ask about these things, but to get at the subjective truths here, I thought it best to use focus groups in which students would talk to one another with a researcher moderating the discussion and guiding it using predetermined questions. You will learn more about focus groups in Chapter 18 .

Researchers spoke with small groups of students during the last class of the semester. They prompted people to talk about aspects of the textbook they liked and didn’t like, compare it to textbooks from other classes, describe how they used it, and so forth. It was this focus on  understanding and subjective experience that brought us into the interpretivist paradigm. Alongside other researchers, I created the focus group questions but encouraged researchers who moderated the focus groups to allow the conversation to flow organically.

We originally started out with the assumption, for which there is support in the literature, that students would be angry with the high-cost textbook that we used prior to the free one, and this cost shock might play a role in students’ negative attitudes about research. But unlike the hypotheses in positivism, these are merely a place to start and are open to revision throughout the research process. This is because the researchers are not the experts, the participants are! Just like your clients are the experts on their lives, so were the students in my study. Our job as researchers was to create a group in which they would reveal their informed thoughts about the issue, coming to consensus around a few key themes.

qualitative meaning of research

When we initially analyzed the focus groups, we uncovered themes that seemed to fit the data. But the overall picture was murky. How were themes related to each other? And how could we distill these themes and relationships into something meaningful? We went back to the data again. We could do this because there isn’t one truth, as in positivism, but multiple truths and multiple ways of interpreting the data. When we looked again, we focused on some of the effects of having a textbook customized to the course. It was that customization process that helped make the language more approachable, engaging, and relevant to social work practice.

Ultimately, our data revealed differences in how students perceived a free textbook versus a free textbook that is customized to the class. When we went to interpret this finding, the remix  hypothesis of open textbook was helpful in understanding that relationship. It states that the more faculty incorporate editing and creating into the course, the better student learning will be. Our study helped flesh out that theory by discussing the customization process and how students made sense of a customized resource.

In this way, theoretical analysis operates differently in interpretivist research. While positivist research tests existing theories, interpretivist research creates theories based on the stories of research participants. However, it is difficult to say if this theory was totally emergent in the dataset or if my prior knowledge of the remix hypothesis influenced my thinking about the data. Interpretivist researchers are encouraged to put a box around their prior experiences and beliefs, acknowledging them, but trying to approach the data with fresh eyes. Interpretivists know that this is never perfectly possible, though, as we are always influenced by our previous experiences when interpreting data and conducting scientific research projects.

Thinking like an activist: Critical

Although adding focus groups helped ease my concern about reducing student learning down to just final grades by providing a more rich set of conversations to analyze. However, my role as researcher and “expert” was still an important part of the analysis. As someone who has been out of school for a while, and indeed has taught this course for years, I have lost touch with what it is like to be a student taking research methods for the first time. How could I accurately interpret or understand what students were saying? Perhaps I would overlook things that reflected poorly on my teaching or my book. I brought other faculty researchers on board to help me analyze the data, but this still didn’t feel like enough.

By luck, an undergraduate student approached me about wanting to work together on a research project. I asked her if she would like to collaborate on evaluating the textbook with me. Over the next year, she assisted me with conceptualizing the project, creating research questions, as well as conducting and analyzing the focus groups. Not only would she provide an “insider” perspective on coding the data, steeped in her lived experience as a student, but she would serve as a check on my power through the process.

Including people from the group you are measuring as part of your research team is a common component of critical research. Ultimately, critical theorists would find my study to be inadequate in many ways. I still developed the research question, created the intervention, and wrote up the results for publication, which privileges my voice and role as “expert.” Instead, critical theorists would emphasize the role of students (community members) in identifying research questions, choosing the best intervention to used, and so forth. But collaborating with students as part of a research team did address some of the power imbalances in the research process.

Critical research projects also aim to have an impact on the people and systems involved in research. No students or researchers had profound personal realizations as a result of my study, nor did it lessen the impact of oppressive structures in society. I can claim some small victory that my department switched to using my textbook after the study was complete (changing a system), though this was likely the result of factors other than the study (my advocacy for open textbooks).

Social work research is almost always designed to create change for people or systems. To that end, every social work project is at least somewhat critical. However, the additional steps of conducting research with people rather than on people reveal a depth to the critical paradigm. By bringing students on board the research team, study had student perspectives represented in conceptualization, data collection, and analysis. That said, there was much to critique about this study from a critical perspective. I retained a lot of the power in the research process, and students did not have the ability to determine the research question or purpose of the project. For example, students might likely have said that textbook costs and the quality of their research methods textbook were less important than student debt, racism, or other potential issues experienced by students in my class. Instead of a ground-up research process based in community engagement, my research included some important participation by students on project created and led by faculty.

Conceptualization is an iterative process

I hope this conversation was useful in applying paradigms to a research project. While my example discusses education research, the same would apply for social work research about social welfare programs, clinical interventions, or other topics. Paradigm and theory are covered at the beginning of the conceptualization of your project because these assumptions will structure the rest of your project. Each of the research steps that occur after this chapter (e.g., forming a question, choosing a design) rely upon philosophical and theoretical assumptions. As you continue conceptualizing your project over the next few weeks, you may find yourself shifting between paradigms. That is normal, as conceptualization is not a linear process. As you move through the next steps of conceptualizing and designing a project, you’ll find philosophies and theories that best match how you want to study your topic.

Viewing theoretical and empirical arguments through this lens is one of the true gifts of the social work approach to research. The multi-paradigmatic perspective is a hallmark of social work research and one that helps us contribute something unique on research teams and in practice.

  • Multi-paradigmatic research is a distinguishing hallmark of social work research. Understanding the limitations and strengths of each paradigm will help you justify your research approach and strategically choose elements from one or more paradigms to answer your question.
  • Paradigmatic assumptions help you understand the “blind spots” in your research project and how to adjust and address these areas. Keep in mind, it is not necessary to address all of your blind spots, as all projects have limitations.
  • Sketch out which paradigm applies best to your project. Second, building on your answer to the exercise in section 7.3, identify how the theory you chose and the paradigm in which you find yourself are consistent or are in conflict with one another. For example, if you are using systems theory in a positivist framework, you might talk about how they both rely on a deterministic approach to human behavior with a focus on the status-quo and social order.
  • Define and provide an example of an idiographic causal explanation
  • Differentiate between idiographic and nomothetic causal relationships
  • Link idiographic and nomothetic causal relationships with the process of theory building and theory testing
  • Describe how idiographic and nomothetic causal explanations can be complementary

As we transition away from positivism, it is important to highlight the assumptions it makes about the scientific process–the hypothetico-deductive method, sometimes referred to as the research circle.

The hypothetico-deductive method

The primary way that researchers in the positivist paradigm use theories is sometimes called the hypothetico-deductive method (although this term is much more likely to be used by philosophers of science than by scientists themselves). Researchers choose an existing theory. Then, they make a prediction about some new phenomenon that should be observed if the theory is correct. Again, this prediction is called a hypothesis. The researchers then conduct an empirical study to test the hypothesis. Finally, they reevaluate the theory in light of the new results and revise it if necessary.

This process is usually conceptualized as a cycle because the researchers can then derive a new hypothesis from the revised theory, conduct a new empirical study to test the hypothesis, and so on. As Figure 8.8 shows, this approach meshes nicely with the process of conducting a research project—creating a more detailed model of “theoretically motivated” or “theory-driven” research. Together, they form a model of theoretically motivated research. 

qualitative meaning of research

Keep in mind the hypothetico-deductive method is only one way of using social theory to inform social science research. It starts with describing one or more existing theories, deriving a hypothesis from one of those theories, testing your hypothesis in a new study, and finally reevaluating the theory based on the results data analyses. This format works well when there is an existing theory that addresses the research question—especially if the resulting hypothesis is surprising or conflicts with a hypothesis derived from a different theory.

But what if your research question is more interpretive? What if it is less about theory-testing and more about theory-building? This is what our next chapter covers: the process of inductively deriving theory from people’s stories and experiences. This process looks different than that depicted in Figure 8.8. It still starts with your research question and answering that question by conducting a research study. But instead of testing a hypothesis you created based on a theory, you will create a theory of your own that explain the data you collected. This format works well for qualitative research questions and for research questions that existing theories do not address.

Inductive reasoning is most commonly found in studies using qualitative methods, such as focus groups and interviews. Because inductive reasoning involves the creation of a new theory, researchers need very nuanced data on how the key concepts in their working question operate in the real world. Qualitative data is often drawn from lengthy interactions and observations with the individuals and phenomena under examination. For this reason, inductive reasoning is most often associated with qualitative methods, though it is used in both quantitative and qualitative research.

qualitative meaning of research

Whose truth does science establish?

Social work is concerned with the “isms” of oppression (ableism, ageism, cissexism, classism, heterosexism, racism, sexism, etc.), and so our approach to science must reconcile its history as both a tool of oppression and its exclusion of oppressed groups. Science grew out of the Enlightenment, a philosophical movement which applied reason and empirical analysis to understanding the world. While the Enlightenment brought forth tremendous achievements, the critiques of Marxian, feminist, and other critical theorists complicated the Enlightenment understanding of science. For this section, I will focus on feminist critiques of science, building upon an entry in the Stanford Encyclopedia of Philosophy (Crasnow, 2020). [16]

In its original formulation, science was an individualistic endeavor. As we learned in Chapter 1 , a basic statement of the scientific method is that a researcher studies existing theories on a topic, formulates a hypothesis about what might be true, and either confirms or disconfirms their hypothesis through experiment and rigorous observation. Over time, our theories become more accurate in their predictions and more comprehensive in their conclusions. Scientists put aside their preconceptions, look at the data, and build their theories based on objective rationality.

Yet, this cannot be perfectly true. Scientists are human, after all. As a profession historically dominated by white men, scientists have dismissed women and other minorities as being psychologically unfit for the scientific profession. While attitudes have improved, science, technology, engineering, mathematics (STEM) and related fields remain dominated by white men (Grogan, 2019). [17] Biases can persist in social work theory and research when social scientists do not have similar experiences to the populations they study.

Gender bias can influence the research questions scientists choose to answer. Feminist critiques of medical science drew attention to women’s health issues, spurring research and changing standards of care. The focus on domestic violence in the empirical literature can also be seen as a result of feminist critique. Thus, critical theory helps us critique what is on the agenda for science. If science is to answer important questions, it must speak to the concerns of all people. Through the democratization in access to scientific knowledge and the means to produce it, science becomes a sister process of social development and social justice.

The goal of a diverse and participatory scientific community lies in contrast to much of what we understand to be “proper” scientific knowledge. Many of the older, classic social science theories were developed based on research which observed males or from university students in the United States or other Western nations. How these observations were made, what questions were asked, and how the data were interpreted were shaped by the same oppressive forces that existed in broader society, a process that continues into the present. In psychology, the concept of hysteria or hysterical women was believed to be caused by a wandering womb (Tasca et al., 2012). [18] Even today, there are gender biases in diagnoses of histrionic personality disorder and racial biases in psychotic disorders (Klonsky et al., 2002) [19] because the theories underlying them were created in a sexist and racist culture. In these ways, science can reinforce the truth of the white Western male perspective.

Finally, it is important to note that social science research is often conducted on populations rather than with populations. Historically, this has often meant Western men traveling to other countries and seeking to understand other cultures through a Western lens. Lacking cultural humility and failing to engage stakeholders, ethnocentric research of this sort has led to the view of non-Western cultures as inferior. Moreover, the use of these populations as research subjects rather than co-equal participants in the research process privileges the researcher’s knowledge over that from other groups or cultures. Researchers working with indigenous cultures, in particular, had a destructive habit of conducting research for a short time and then leaving, without regard for the impact their study had on the population. These critiques of Western science aim to decolonize social science and dismantle the racist ideas the oppress indigenous and non-Western peoples through research (Smith, 2013). [20]

The central concept in feminist, anti-racist, and decolonization critiques (among other critical frames) is epistemic injustice. Epistemic injustice happens when someone is treated unfairly in their capacity to know something or describe their experience of the world. As described by Fricker (2011), [21] the injustice emerges from the dismissal of knowledge from oppressed groups, discrimination against oppressed groups in scientific communities, and the resulting gap between what scientists can make sense of from their experience and the experiences of people with less power who have lived experience of the topic. We recommend this video from Edinburgh Law School which applies epistemic injustice to studying public health emergencies, disabilities, and refugee services .

The letters IV on the left side with an arrow pointing to the letters DV on the right

Positivism relies on nomothetic causality, or the idea that “one event, behavior, or belief will result in the occurrence of another, subsequent event, behavior, or belief.” Then, we described one kind of causality: a simple cause-and-effect relationship supported by existing theory and research on the topic, also known as a nomothetic causal relationship. But what if there is not a lot of literature on your topic? What if your question is more exploratory than explanatory? Then, you need a different kind of causal explanation, one that accounts for the complexity of human interactions.

How can we build causal relationships if we are just describing or exploring a topic? Recall the definitions of exploratory research , descriptive research , and explanatory research from Chapter 2. Wouldn’t we need to do explanatory research to build any kind of causal explanation? Explanatory research attempts to establish nomothetic causal relationships: an independent variable is demonstrated to cause change in a dependent variable. Exploratory and descriptive qualitative research contains some causal relationships, but they are actually descriptions of the causal relationships established by the study participants.

What do idiographic causal explanations look like?

An idiographic causal relationship   tries to identify the many, interrelated causes that account for the phenomenon the researcher is investigating. So, if idiographic causal explanations do not look like Figure 8.5, 8.6, or 8.7 what do they look like? Instead of saying “x causes y,” your participants will describe their experiences with “x,” which they will tell you was caused and influenced by a variety of other factors, as interpreted through their unique perspective, time, and environment. As we stated before, idiographic causal explanations are messy. Your job as a social science researcher is to accurately describe the patterns in what your participants tell you.

Let’s think about this using an example. If I asked you why you decided to become a social worker, what might you say? For me, I would say that I wanted to be a mental health clinician since I was in high school. I was interested in how people thought, and I was privileged enough to have psychology courses at my local high school. I thought I wanted to be a psychologist, but at my second internship in my undergraduate program, my supervisors advised me to become a social worker because the license provided greater authority for independent practice and flexibility for career change. Once I found out social workers were like psychologists who also raised trouble about social justice, I was hooked.

That’s not a simple explanation at all! But it’s definitely a causal explanation. It is my individual, subjective truth of a complex process. If we were to ask multiple social workers the same question, we might find out that many social workers begin their careers based on factors like personal experience with a disability or social injustice, positive experiences with social workers, or a desire to help others. No one factor is the “most important factor,” like with nomothetic causal relationships. Instead, a complex web of factors, contingent on context, emerge when you interpret what people tell you about their lives.

Understanding “why?”

In creating an idiographic explanation, you are still asking “why?” But the answer is going to be more complex. Those complexities are described in Table 8.1 as well as this short video comparing nomothetic and idiographic relationships .

Table 8.1: Comparing nomothetic and idiographic causal relationships
Nomothetic causal relationships Idiographic causal relationships
Paradigm Positivist Interpretivist
Purpose of research Prediction & generalization Understanding & particularity
Reasoning Deductive Inductive
Purpose of research Explanatory Exploratory or descriptive
Research methods Quantitative Qualitative
Causality Simple: cause and effect Complex: context-dependent, sometimes circular or contradictory
Role of theory Theory testing Theory building

Remember our question from the last section, “Are you trying to generalize or nah?” If you answered nah (or no, like a normal person), you are trying to establish an idiographic causal explanation. The purpose of that explanation isn’t to predict the future or generalize to larger populations, but to describe the here-and-now as it is experienced by individuals within small groups and communities. Idiographic explanations are focused less on what is generally experienced by all people but more on the particularities of what specific individuals in a unique time and place experience.

Researchers seeking idiographic causal relationships are not trying to generalize or predict, so they have no need to reduce phenomena to mathematics. In fact, only examining things that can be counted can rob a causal relationship of its meaning and context. Instead, the goal of idiographic causal relationships is understanding, rather than prediction. Idiographic causal relationships are formed by interpreting people’s stories and experiences. Usually, these are expressed through words. Not all qualitative studies use word data, as some can use interpretations of visual or performance art. However, the vast majority of qualitative studies do use word data, like the transcripts from interviews and focus groups or documents like journal entries or meeting notes. Your participants are the experts on their lives—much like in social work practice—and as in practice, people’s experiences are embedded in their cultural, historical, and environmental context.

Idiographic causal explanations are powerful because they can describe the complicated and interconnected nature of human life. Nomothetic causal explanations, by comparison, are simplistic. Think about if someone asked you why you wanted to be a social worker. Your story might include a couple of vignettes from your education and early employment. It might include personal experience with the social welfare system or family traditions. Maybe you decided on a whim to enroll in a social work course during your graduate program. The impact of each of these events on your career is unique to you.

Idiographic causal explanations are concerned with individual stories, their idiosyncrasies, and the patterns that emerge when you collect and analyze multiple people’s stories. This is the inductive reasoning we discussed at the beginning of this chapter. Often, idiographic causal explanations begin by collecting a lot of qualitative data, whether though interviews, focus groups, or looking at available documents or cultural artifacts. Next, the researcher looks for patterns in the data and arrives at a tentative theory for how the key ideas in people’s stories are causally related.

Unlike nomothetic causal relationships, there are no formal criteria (e.g., covariation) for establishing causality in idiographic causal relationships. In fact, some criteria like temporality and nonspuriousness may be violated. For example, if an adolescent client says, “It’s hard for me to tell whether my depression began before my drinking, but both got worse when I was expelled from my first high school,” they are recognizing that it may not so simple that one thing causes another. Sometimes, there is a reciprocal relationship where one variable (depression) impacts another (alcohol abuse), which then feeds back into the first variable (depression) and into other variables as well (school). Other criteria, such as covariation and plausibility, still make sense, as the relationships you highlight as part of your idiographic causal explanation should still be plausible and its elements should vary together.

Theory building and theory testing

As we learned in the previous section, nomothetic causal explanations are created by researchers applying deductive reasoning to their topic and creating hypotheses using social science theories. Much of what we think of as social science is based on this hypothetico-deductive method, but this leaves out the other half of the equation. Where do theories come from? Are they all just revisions of one another? How do any new ideas enter social science?

Through inductive reasoning and idiographic causal explanations!

Let’s consider a social work example. If you plan to study domestic and sexual violence, you will likely encounter the Power and Control Wheel, also known as the Duluth Model (Figure 8.9). The wheel is a model designed to depict the process of domestic violence. The wheel was developed based on qualitative focus groups conducted by sexual and domestic violence advocates in Duluth, MN. This video explains more about the Duluth Model of domestic abuse.

Power and control wheel indicating the factors like

The Power and Control Wheel is an example of what an idiographic causal relationship looks like. By contrast, look back at the previous section’s Figure 8.5, 8.6, and 8.7 on nomothetic causal relationships between independent and dependent variables. See how much more complex idiographic causal explanations are?! They are complex, but not difficult to understand. At the center of domestic abuse is power and control, and while not every abuser would say that is what they were doing, that is the understanding of the survivors who informed this theoretical model. Their power and control is maintained through a variety of abusive tactics from social isolation to use of privilege to avoid consequences.

What about the role of hypotheses in idiographic causal explanations? In nomothetic causal explanations, researchers create hypotheses using existing theory and then test them for accuracy. Hypotheses in idiographic causality are much more tentative, and are probably best considered as “hunches” about what they think might be true. Importantly, they might indicate the researcher’s prior knowledge and biases before the project begins, but the goal of idiographic research is to let your participants guide you rather than existing social work knowledge. Continuing with our Duluth Model example, advocates likely had some tentative hypotheses about what was important in a relationship with domestic violence. After all, they worked with this population for years prior to the creation of the model. However, it was the stories of the participants in these focus groups that led the Power and Control Wheel explanation for domestic abuse.

As qualitative inquiry unfolds, hypotheses and hunches are likely to emerge and shift as researchers learn from what their participants share. Because the participants are the experts in idiographic causal relationships, a researcher should be open to emerging topics and shift their research questions and hypotheses accordingly. This is in contrast to hypotheses in quantitative research, which remain constant throughout the study and are shown to be true or false.

Over time, as more qualitative studies are done and patterns emerge across different studies and locations, more sophisticated theories emerge that explain phenomena across multiple contexts. Once a theory is developed from qualitative studies, a quantitative researcher can seek to test that theory. For example, a quantitative researcher may hypothesize that men who hold traditional gender roles are more likely to engage in domestic violence. That would make sense based on the Power and Control Wheel model, as the category of “using male privilege” speaks to this relationship. In this way, qualitatively-derived theory can inspire a hypothesis for a quantitative research project, as we will explore in the next section.

Complementary approaches

If idiographic and nomothetic still seem like obscure philosophy terms, let’s consider another example. Imagine you are working for a community-based non-profit agency serving people with disabilities. You are putting together a report to lobby the state government for additional funding for community support programs. As part of that lobbying, you are likely to rely on both nomothetic and idiographic causal relationships.

If you looked at nomothetic causal relationships, you might learn how previous studies have shown that, in general, community-based programs like yours are linked with better health and employment outcomes for people with disabilities. Nomothetic causal explanations seek to establish that community-based programs are better for everyone with disabilities, including people in your community.

If you looked at idiographic causal explanations, you would use stories and experiences of people in community-based programs. These individual stories are full of detail about the lived experience of being in a community-based program. You might use one story from a client in your lobbying campaign, so policymakers can understand the lived experience of what it’s like to be a person with a disability in this program. For example, a client who said “I feel at home when I’m at this agency because they treat me like a family member,” or “this is the agency that helped me get my first paycheck,” can communicate richer, more complex causal relationships.

Neither kind of causal explanation is better than the other. A decision to seek idiographic causal explanations means that you will attempt to explain or describe your phenomenon exhaustively, attending to cultural context and subjective interpretations. A decision to seek nomothetic causal explanations, on the other hand, means that you will try to explain what is true for everyone and predict what will be true in the future. In short, idiographic explanations have greater depth, and nomothetic explanations have greater breadth.

Most importantly, social workers understand the value of both approaches to understanding the social world. A social worker helping a client with substance abuse issues seeks idiographic explanations when they ask about that client’s life story, investigate their unique physical environment, or probe how their family relationships. At the same time, a social worker also uses nomothetic explanations to guide their interventions. Nomothetic explanations may help guide them to minimize risk factors and maximize protective factors or use an evidence-based therapy, relying on knowledge about what in general  helps people with substance abuse issues.

So, which approach speaks to you? Are you interested in learning about (a) a few people’s experiences in a great deal of depth, or (b) a lot of people’s experiences more superficially, while also hoping your findings can be generalized to a greater number of people? The answer to this question will drive your research question and project. These approaches provide different types of information and both types are valuable.

  • Idiographic causal explanations focus on subjectivity, context, and meaning.
  • Idiographic causal explanations are best suited to exploratory research questions and qualitative methods.
  • Idiographic causal explanations are used to create new theories in social science.
  • Explore the literature on the theory you identified in section 8.1.
  • Read about the origins of your theory. Who developed it and from what data?
  • See if you can find a figure like Figure 8.9 in an article or book chapter that depicts the key concepts in your theory and how those concepts are related to one another causally. Write out a short statement on the causal relationships contained in the figure.
  • List the key terms associated with qualitative research questions
  • Distinguish between qualitative and quantitative research questions

Qualitative research questions differ from quantitative research questions. Because qualitative research questions seek to explore or describe phenomena, not provide a neat nomothetic explanation, they are often more general and openly worded. They may include only one concept, though many include more than one. Instead of asking how one variable causes changes in another, we are instead trying to understand the experiences ,  understandings , and  meanings that people have about the concepts in our research question. These keywords often make an appearance in qualitative research questions.

Let’s work through an example from our last section. In Table 9.1, a student asked, “What is the relationship between sexual orientation or gender identity and homelessness for late adolescents in foster care?” In this question, it is pretty clear that the student believes that adolescents in foster care who identify as LGBTQ+ may be at greater risk for homelessness. This is a nomothetic causal relationship—LGBTQ+ status causes changes in homelessness.

However, what if the student were less interested in  predicting  homelessness based on LGBTQ+ status and more interested in  understanding  the stories of foster care youth who identify as LGBTQ+ and may be at risk for homelessness? In that case, the researcher would be building an idiographic causal explanation . The youths whom the researcher interviews may share stories of how their foster families, caseworkers, and others treated them. They may share stories about how they thought of their own sexuality or gender identity and how it changed over time. They may have different ideas about what it means to transition out of foster care.

qualitative meaning of research

Because qualitative questions usually center on idiographic causal relationships, they look different than quantitative questions. Table 9.3 below takes the final research questions from Table 9.1 and adapts them for qualitative research. The guidelines for research questions previously described in this chapter still apply, but there are some new elements to qualitative research questions that are not present in quantitative questions.

  • Qualitative research questions often ask about lived experience, personal experience, understanding, meaning, and stories.
  • Qualitative research questions may be more general and less specific.
  • Qualitative research questions may also contain only one variable, rather than asking about relationships between multiple variables.
Table 9.3 Quantitative vs. qualitative research questions
How does witnessing domestic violence impact a child’s romantic relationships in adulthood? How do people who witness domestic violence understand its effects on their current relationships?
What is the relationship between sexual orientation or gender identity and homelessness for late adolescents in foster care? What is the experience of identifying as LGBTQ+ in the foster care system?
How does income inequality affect ambivalence in high-density urban areas? What does racial ambivalence mean to residents of an urban neighborhood with high income inequality?
How does race impact rates of mental health diagnosis for children in foster care? How do African-Americans experience seeking help for mental health concerns?

Qualitative research questions have one final feature that distinguishes them from quantitative research questions: they can change over the course of a study. Qualitative research is a reflexive process, one in which the researcher adapts their approach based on what participants say and do. The researcher must constantly evaluate whether their question is important and relevant to the participants. As the researcher gains information from participants, it is normal for the focus of the inquiry to shift.

For example, a qualitative researcher may want to study how a new truancy rule impacts youth at risk of expulsion. However, after interviewing some of the youth in their community, a researcher might find that the rule is actually irrelevant to their behavior and thoughts. Instead, their participants will direct the discussion to their frustration with the school administrators or the lack of job opportunities in the area. This is a natural part of qualitative research, and it is normal for research questions and hypothesis to evolve based on information gleaned from participants.

However, this reflexivity and openness unacceptable in quantitative research for good reasons. Researchers using quantitative methods are testing a hypothesis, and if they could revise that hypothesis to match what they found, they could never be wrong! Indeed, an important component of open science and reproducability is the preregistration of a researcher’s hypotheses and data analysis plan in a central repository that can be verified and replicated by reviewers and other researchers. This interactive graphic from 538 shows how an unscrupulous research could come up with a hypothesis and theoretical explanation  after collecting data by hunting for a combination of factors that results in a statistically significant relationship. This is an excellent example of how the positivist assumptions behind quantitative research and intepretivist assumptions behind qualitative research result in different approaches to social science.

  • Qualitative research questions often contain words or phrases like “lived experience,” “personal experience,” “understanding,” “meaning,” and “stories.”
  • Qualitative research questions can change and evolve over the course of the study.
  • Using the guidance in this chapter, write a qualitative research question. You may want to use some of the keywords mentioned above.
  • Kivuna, C. & Kuyini, A. B. (2017). Understanding and applying research paradigms in educational contexts. International Journal of Higher Education, 6 (5), 26-41. https://eric.ed.gov/?id=EJ1154775 ↵
  • Kuhn, T. (1962). The structure of scientific revolutions . Chicago: University of Chicago Press. ↵
  • Fleuridas, C., & Krafcik, D. (2019). Beyond four forces: The evolution of psychotherapy. Sage Open ,  9 (1), 2158244018824492. ↵
  • Shneider, A. M. (2009). Four stages of a scientific discipline; four types of scientist. Trends in Biochemical Sciences 34 (5), 217-233. https://doi.org/10.1016/j.tibs.2009.02.00 ↵
  • Burrell, G. & Morgan, G. (1979). Sociological paradigms and organizational analysis . Routledge. Guba, E. (ed.) (1990). The paradigm dialog . SAGE. ↵
  • Routledge. Guba, E. (ed.) (1990). The paradigm dialog . SAGE. ↵
  • Burrell, G. & Morgan, G. (1979). Sociological paradigms and organizational analysis . Here is a summary of Burrell & Morgan from Babson College , and our classification collapses radical humanism and radical structuralism into the critical paradigm, following Guba and Lincoln's three-paradigm framework. We feel this approach is more parsimonious and easier for students to understand on an introductory level. ↵
  • For more about how the meanings of hand gestures vary by region, you might read the following blog entry: Wong, W. (2007). The top 10 hand gestures you’d better get right . Retrieved from: http://www.languagetrainers.co.uk/blog/2007/09/24/top-10-hand-gestures ↵
  • Rosario, M., Schrimshaw, E. W., Hunter, J., & Levy-Warren, A. (2009). The coming-out process of young lesbian and bisexual women: Are there butch/femme differences in sexual identity development?. Archives of sexual behavior ,  38 (1), 34-49. ↵
  • Calhoun, C., Gerteis, J., Moody, J., Pfaff, S., & Virk, I. (Eds.). (2007). Classical sociological theory  (2nd ed.). Malden, MA: Blackwell. ↵
  • Fraser, N. (1989).  Unruly practices: Power, discourse, and gender in contemporary social theory . Minneapolis, MN: University of Minnesota Press. ↵
  • Here are links to two HBSE open textbooks, if you are unfamiliar with social work theories and would like more background. https://uark.pressbooks.pub/hbse1/ and https://uark.pressbooks.pub/humanbehaviorandthesocialenvironment2/ ↵
  • Box, G. E. P.. (1976). Science and statistics. Journal of the American Statistical Association, 71 (356), 791. ↵
  • Heineman-Pieper, J., Tyson, K., & Pieper, M. H. (2002). Doing good science without sacrificing good values: Why the heuristic paradigm is the best choice for social work.  Families in Society ,  83 (1), 15-28. ↵
  • Crasnow, S. (2020). Feminist perspectives on science. In E. N. Zalta (ed.), The Stanford Encyclopedia of Philosophy (Winter 2020 Edition). Retrieved from: https://plato.stanford.edu/entries/feminist-science/ ↵
  • Grogan, K.E. (2019) How the entire scientific community can confront gender bias in the workplace. Nature Ecology & Evolution, 3 ,  3–6. doi:10.1038/s41559-018-0747-4 ↵
  • Tasca, C., Rapetti, M., Carta, M. G., & Fadda, B. (2012). Women and hysteria in the history of mental health. Clinical practice and epidemiology in mental health: Clinical practice & epidemiology in mental health ,  8 , 110-119. ↵
  • Klonsky, E. D., Jane, J. S., Turkheimer, E., & Oltmanns, T. F. (2002). Gender role and personality disorders.  Journal of personality disorders ,  16 (5), 464-476. ↵
  • Smith, L. T. (2013). Decolonizing methodologies: Research and indigenous peoples . Zed Books Ltd. ↵
  • Fricker, M. (2011). Epistemic injustice: Power and the ethics of knowing . Oxford University Press. ↵

The highest level of measurement. Denoted by mutually exclusive categories, a hierarchy (order), values can be added, subtracted, multiplied, and divided, and the presence of an absolute zero.

a paradigm based on the idea that social context and interaction frame our realities

a paradigm in social science research focused on power, inequality, and social change

a research paradigm that suspends questions of philosophical ‘truth’ and focuses more on how different philosophies, theories, and methods can be used strategically to resolve a problem or question within the researcher's unique context

A cyclical process of theory development, starting with an observed phenomenon, then developing or using a theory to make a specific prediction of what should happen if that theory is correct, testing that prediction, refining the theory in light of the findings, and using that refined theory to develop new hypotheses, and so on.

when someone is treated unfairly in their capacity to know something or describe their experience of the world

conducted during the early stages of a project, usually when a researcher wants to test the feasibility of conducting a more extensive study or if the topic has not been studied in the past

research that describes or defines a particular phenomenon

explains why particular phenomena work in the way that they do; answers “why” questions

attempts to explain or describe your phenomenon exhaustively, based on the subjective understandings of your participants

"Assuming that the null hypothesis is true and the study is repeated an infinite number times by drawing random samples from the same populations(s), less than 5% of these results will be more extreme than the current result" (Cassidy et al., 2019, p. 233).

Scientific Inquiry in Social Work (2nd Edition) Copyright © 2020 by Matthew DeCarlo, Cory Cummings, and Kate Agnelli is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License , except where otherwise noted.

Share This Book

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
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed, qualitative study, affiliations.

  • 1 University of Nebraska Medical Center
  • 2 GDB Research and Statistical Consulting
  • 3 GDB Research and Statistical Consulting/McLaren Macomb Hospital
  • PMID: 29262162
  • Bookshelf ID: NBK470395

Qualitative research is a type of research that explores and provides deeper insights into real-world problems. Instead of collecting numerical data points or intervening or introducing treatments just like in quantitative research, qualitative research helps generate hypothenar to further investigate and understand quantitative data. Qualitative research gathers participants' experiences, perceptions, and behavior. It answers the hows and whys instead of how many or how much. It could be structured as a standalone study, purely relying on qualitative data, or part of mixed-methods research that combines qualitative and quantitative data. This review introduces the readers to some basic concepts, definitions, terminology, and applications of qualitative research.

Qualitative research, at its core, asks open-ended questions whose answers are not easily put into numbers, such as "how" and "why." Due to the open-ended nature of the research questions, qualitative research design is often not linear like quantitative design. One of the strengths of qualitative research is its ability to explain processes and patterns of human behavior that can be difficult to quantify. Phenomena such as experiences, attitudes, and behaviors can be complex to capture accurately and quantitatively. In contrast, a qualitative approach allows participants themselves to explain how, why, or what they were thinking, feeling, and experiencing at a particular time or during an event of interest. Quantifying qualitative data certainly is possible, but at its core, qualitative data is looking for themes and patterns that can be difficult to quantify, and it is essential to ensure that the context and narrative of qualitative work are not lost by trying to quantify something that is not meant to be quantified.

However, while qualitative research is sometimes placed in opposition to quantitative research, where they are necessarily opposites and therefore "compete" against each other and the philosophical paradigms associated with each other, qualitative and quantitative work are neither necessarily opposites, nor are they incompatible. While qualitative and quantitative approaches are different, they are not necessarily opposites and certainly not mutually exclusive. For instance, qualitative research can help expand and deepen understanding of data or results obtained from quantitative analysis. For example, say a quantitative analysis has determined a correlation between length of stay and level of patient satisfaction, but why does this correlation exist? This dual-focus scenario shows one way in which qualitative and quantitative research could be integrated.

Copyright © 2024, StatPearls Publishing LLC.

PubMed Disclaimer

Conflict of interest statement

Disclosure: Steven Tenny declares no relevant financial relationships with ineligible companies.

Disclosure: Janelle Brannan declares no relevant financial relationships with ineligible companies.

Disclosure: Grace Brannan declares no relevant financial relationships with ineligible companies.

  • Introduction
  • Issues of Concern
  • Clinical Significance
  • Enhancing Healthcare Team Outcomes
  • Review Questions

Similar articles

  • Folic acid supplementation and malaria susceptibility and severity among people taking antifolate antimalarial drugs in endemic areas. Crider K, Williams J, Qi YP, Gutman J, Yeung L, Mai C, Finkelstain J, Mehta S, Pons-Duran C, Menéndez C, Moraleda C, Rogers L, Daniels K, Green P. Crider K, et al. Cochrane Database Syst Rev. 2022 Feb 1;2(2022):CD014217. doi: 10.1002/14651858.CD014217. Cochrane Database Syst Rev. 2022. PMID: 36321557 Free PMC article.
  • Macromolecular crowding: chemistry and physics meet biology (Ascona, Switzerland, 10-14 June 2012). Foffi G, Pastore A, Piazza F, Temussi PA. Foffi G, et al. Phys Biol. 2013 Aug;10(4):040301. doi: 10.1088/1478-3975/10/4/040301. Epub 2013 Aug 2. Phys Biol. 2013. PMID: 23912807
  • The future of Cochrane Neonatal. Soll RF, Ovelman C, McGuire W. Soll RF, et al. Early Hum Dev. 2020 Nov;150:105191. doi: 10.1016/j.earlhumdev.2020.105191. Epub 2020 Sep 12. Early Hum Dev. 2020. PMID: 33036834
  • Invited review: Qualitative research in dairy science-A narrative review. Ritter C, Koralesky KE, Saraceni J, Roche S, Vaarst M, Kelton D. Ritter C, et al. J Dairy Sci. 2023 Sep;106(9):5880-5895. doi: 10.3168/jds.2022-23125. Epub 2023 Jul 18. J Dairy Sci. 2023. PMID: 37474366 Review.
  • Participation in environmental enhancement and conservation activities for health and well-being in adults: a review of quantitative and qualitative evidence. Husk K, Lovell R, Cooper C, Stahl-Timmins W, Garside R. Husk K, et al. Cochrane Database Syst Rev. 2016 May 21;2016(5):CD010351. doi: 10.1002/14651858.CD010351.pub2. Cochrane Database Syst Rev. 2016. PMID: 27207731 Free PMC article. Review.
  • Moser A, Korstjens I. Series: Practical guidance to qualitative research. Part 1: Introduction. Eur J Gen Pract. 2017 Dec;23(1):271-273. - PMC - PubMed
  • Cleland JA. The qualitative orientation in medical education research. Korean J Med Educ. 2017 Jun;29(2):61-71. - PMC - PubMed
  • Foley G, Timonen V. Using Grounded Theory Method to Capture and Analyze Health Care Experiences. Health Serv Res. 2015 Aug;50(4):1195-210. - PMC - PubMed
  • Devers KJ. How will we know "good" qualitative research when we see it? Beginning the dialogue in health services research. Health Serv Res. 1999 Dec;34(5 Pt 2):1153-88. - PMC - PubMed
  • Huston P, Rowan M. Qualitative studies. Their role in medical research. Can Fam Physician. 1998 Nov;44:2453-8. - PMC - PubMed

Publication types

  • Search in PubMed
  • Search in MeSH
  • Add to Search

Related information

  • Cited in Books

LinkOut - more resources

Full text sources.

  • NCBI Bookshelf

book cover photo

  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

Have a language expert improve your writing

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

  • Knowledge Base

Methodology

  • Qualitative vs. Quantitative Research | Differences, Examples & Methods

Qualitative vs. Quantitative Research | Differences, Examples & Methods

Published on April 12, 2019 by Raimo Streefkerk . Revised on June 22, 2023.

When collecting and analyzing data, quantitative research deals with numbers and statistics, while qualitative research deals with words and meanings. Both are important for gaining different kinds of knowledge.

Common quantitative methods include experiments, observations recorded as numbers, and surveys with closed-ended questions.

Quantitative research is at risk for research biases including information bias , omitted variable bias , sampling bias , or selection bias . Qualitative research Qualitative research is expressed in words . It is used to understand concepts, thoughts or experiences. This type of research enables you to gather in-depth insights on topics that are not well understood.

Common qualitative methods include interviews with open-ended questions, observations described in words, and literature reviews that explore concepts and theories.

Table of contents

The differences between quantitative and qualitative research, data collection methods, when to use qualitative vs. quantitative research, how to analyze qualitative and quantitative data, other interesting articles, frequently asked questions about qualitative and quantitative research.

Quantitative and qualitative research use different research methods to collect and analyze data, and they allow you to answer different kinds of research questions.

Qualitative vs. quantitative research

Quantitative and qualitative data can be collected using various methods. It is important to use a data collection method that will help answer your research question(s).

Many data collection methods can be either qualitative or quantitative. For example, in surveys, observational studies or case studies , your data can be represented as numbers (e.g., using rating scales or counting frequencies) or as words (e.g., with open-ended questions or descriptions of what you observe).

However, some methods are more commonly used in one type or the other.

Quantitative data collection methods

  • Surveys :  List of closed or multiple choice questions that is distributed to a sample (online, in person, or over the phone).
  • Experiments : Situation in which different types of variables are controlled and manipulated to establish cause-and-effect relationships.
  • Observations : Observing subjects in a natural environment where variables can’t be controlled.

Qualitative data collection methods

  • Interviews : Asking open-ended questions verbally to respondents.
  • Focus groups : Discussion among a group of people about a topic to gather opinions that can be used for further research.
  • Ethnography : Participating in a community or organization for an extended period of time to closely observe culture and behavior.
  • Literature review : Survey of published works by other authors.

A rule of thumb for deciding whether to use qualitative or quantitative data is:

  • Use quantitative research if you want to confirm or test something (a theory or hypothesis )
  • Use qualitative research if you want to understand something (concepts, thoughts, experiences)

For most research topics you can choose a qualitative, quantitative or mixed methods approach . Which type you choose depends on, among other things, whether you’re taking an inductive vs. deductive research approach ; your research question(s) ; whether you’re doing experimental , correlational , or descriptive research ; and practical considerations such as time, money, availability of data, and access to respondents.

Quantitative research approach

You survey 300 students at your university and ask them questions such as: “on a scale from 1-5, how satisfied are your with your professors?”

You can perform statistical analysis on the data and draw conclusions such as: “on average students rated their professors 4.4”.

Qualitative research approach

You conduct in-depth interviews with 15 students and ask them open-ended questions such as: “How satisfied are you with your studies?”, “What is the most positive aspect of your study program?” and “What can be done to improve the study program?”

Based on the answers you get you can ask follow-up questions to clarify things. You transcribe all interviews using transcription software and try to find commonalities and patterns.

Mixed methods approach

You conduct interviews to find out how satisfied students are with their studies. Through open-ended questions you learn things you never thought about before and gain new insights. Later, you use a survey to test these insights on a larger scale.

It’s also possible to start with a survey to find out the overall trends, followed by interviews to better understand the reasons behind the trends.

Qualitative or quantitative data by itself can’t prove or demonstrate anything, but has to be analyzed to show its meaning in relation to the research questions. The method of analysis differs for each type of data.

Analyzing quantitative data

Quantitative data is based on numbers. Simple math or more advanced statistical analysis is used to discover commonalities or patterns in the data. The results are often reported in graphs and tables.

Applications such as Excel, SPSS, or R can be used to calculate things like:

  • Average scores ( means )
  • The number of times a particular answer was given
  • The correlation or causation between two or more variables
  • The reliability and validity of the results

Analyzing qualitative data

Qualitative data is more difficult to analyze than quantitative data. It consists of text, images or videos instead of numbers.

Some common approaches to analyzing qualitative data include:

  • Qualitative content analysis : Tracking the occurrence, position and meaning of words or phrases
  • Thematic analysis : Closely examining the data to identify the main themes and patterns
  • Discourse analysis : Studying how communication works in social contexts

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.

  • Chi square goodness of fit test
  • Degrees of freedom
  • Null hypothesis
  • Discourse analysis
  • Control groups
  • Mixed methods research
  • Non-probability sampling
  • Quantitative research
  • Inclusion and exclusion criteria

Research bias

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

Quantitative research deals with numbers and statistics, while qualitative research deals with words and meanings.

Quantitative methods allow you to systematically measure variables and test hypotheses . Qualitative methods allow you to explore concepts and experiences in more detail.

In mixed methods research , you use both qualitative and quantitative data collection and analysis methods to answer your research question .

The research methods you use depend on the type of data you need to answer your research question .

  • If you want to measure something or test a hypothesis , use quantitative methods . If you want to explore ideas, thoughts and meanings, use qualitative methods .
  • If you want to analyze a large amount of readily-available data, use secondary data. If you want data specific to your purposes with control over how it is generated, collect primary data.
  • If you want to establish cause-and-effect relationships between variables , use experimental methods. If you want to understand the characteristics of a research subject, use descriptive methods.

Data collection is the systematic process by which observations or measurements are gathered in research. It is used in many different contexts by academics, governments, businesses, and other organizations.

There are various approaches to qualitative data analysis , but they all share five steps in common:

  • Prepare and organize your data.
  • Review and explore your data.
  • Develop a data coding system.
  • Assign codes to the data.
  • Identify recurring themes.

The specifics of each step depend on the focus of the analysis. Some common approaches include textual analysis , thematic analysis , and discourse analysis .

A research project is an academic, scientific, or professional undertaking to answer a research question . Research projects can take many forms, such as qualitative or quantitative , descriptive , longitudinal , experimental , or correlational . What kind of research approach you choose will depend on your topic.

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.

Streefkerk, R. (2023, June 22). Qualitative vs. Quantitative Research | Differences, Examples & Methods. Scribbr. Retrieved September 16, 2024, from https://www.scribbr.com/methodology/qualitative-quantitative-research/

Is this article helpful?

Raimo Streefkerk

Raimo Streefkerk

Other students also liked, what is quantitative research | definition, uses & methods, what is qualitative research | methods & examples, mixed methods research | definition, guide & examples, what is your plagiarism score.

  • Open access
  • Published: 16 September 2024

Gaps in communication theory paradigms when conducting implementation science research: qualitative observations from interviews with administrators, implementors, and evaluators of rural health programs

  • Nicole L. Johnson   ORCID: orcid.org/0000-0001-5686-2062 1 , 2 ,
  • Jennifer Van Tiem 1 , 2 ,
  • Erin Balkenende 1 , 3 ,
  • DeShauna Jones 1 , 4 ,
  • Julia E. Friberg 1 , 2 ,
  • Emily E. Chasco 1 , 4 ,
  • Jane Moeckli 1 , 2 ,
  • Kenda S. Steffensmeier 1 , 2 ,
  • Melissa J. A. Steffen 1 , 2 ,
  • Kanika Arora 5 ,
  • Borsika A. Rabin 6 , 7 &
  • Heather Schacht Reisinger 1 , 3 , 4  

Implementation Science volume  19 , Article number:  66 ( 2024 ) Cite this article

3 Altmetric

Metrics details

Communication is considered an inherent element of nearly every implementation strategy. Often it is seen as a means for imparting new information between stakeholders, representing a Transaction orientation to communication. From a Process orientation, communication is more than information-exchange and is acknowledged as being shaped by (and shaping) the individuals involved and their relationships with one another. As the field of Implementation Science (IS) works to strengthen theoretical integration, we encourage an interdisciplinary approach that engages communication theory to develop richer understanding of strategies and determinants of practice.

We interviewed 28 evaluators, 12 implementors, and 12 administrators from 21 Enterprise-Wide Initiatives funded by the Department of Veteran Affairs Office of Rural Health. Semi-structured interviews focused on experiences with implementation and evaluation strategies. We analyzed the interviews using thematic analysis identifying a range of IS constructs. Then we deductively classified those segments based on a Transaction or Process orientation to communication.

We organized findings using the two IS constructs most commonly discussed in interviews: Collaboration and Leadership Buy-in. The majority of segments coded as Collaboration ( n  = 34, 74%) and Leadership Buy-in ( n  = 31, 70%) discussed communication from a Transaction orientation and referred to communication as synonymous with information exchange, which emphasizes the task over the relationships between the individuals performing the tasks. Conversely, when participants discussed Collaboration and Leadership Buy-in from a Process orientation, they acknowledged both constructs as the result of long-term efforts to develop positive relationships based on trust and respect, and emphasized the time costliness of such strategies. Our findings demonstrate that participants who discussed communication from a Process orientation recognized the nuance and complexity of interpersonal interactions, particularly in the context of IS.

Conclusions

Efficient, reliable information exchange is a critical but often overemphasized element of implementation. Practitioners and researchers must recognize and incorporate the larger role of communication in IS. Two suggestions for engaging a Process orientation to communication are to: (a) use interview probes to learn how communication is enacted, and (b) use process-oriented communication theories to develop interventions and evaluation tools.

Peer Review reports

Contributions to the literature

Communication is a vital part of implementation. Yet, predominant discussions about implementation strategies are limited to a Transactional orientation. Conversely, the Process orientation to communication acknowledges the multiple moving elements in an implementation context that influences collaboration and leadership buy-in.

Exemplars of interview segments about communication engaging a Process orientation were identified to demonstrate ways interviewers can probe to gain a deeper understanding of communication as a process.

We provide examples and suggestions for qualitatively examining communication processes to better understand the impact of implementation strategies.

Several theories with a Process orientation are identified for consideration in future research and implementation planning and evaluation.

Most implementation strategies include a communication component, particularly when evidence-based interventions are introduced and promoted throughout an organization. When implementing new programming, it is common to consider communication as simply a means through which information is imparted [ 1 , 2 ]. Implementation Science (IS) researchers have an imperative to understand the role of communication as more than a means for information exchange [ 3 ]. Yet, even as a means for information exchange, Manojlovich and colleagues recognized the lack of attention on communication in implementation research [ 1 ].

Broadly, the study of communication focuses on how messages are used to generate meanings [ 4 ], and provides perspective for moving beyond an emphasis on information exchange, thus moving beyond the task dimension and recognizing the value of the relational dimension. Despite its relatively young development both academically and professionally, the communication discipline offers valuable insight to IS research [ 5 ]. There are two predominant ways to characterize communication: (1) communication as Transaction, and (2) communication as Process. When communication is viewed as a Transaction, it is discussed as a linear one-way flow of information [ 3 ]. The materiality – the element of substantive value – of communication is found in accurate, efficient information transfer, thus putting emphasis on the task dimension and channel (e.g., phone, handout) through which information is exchanged. When practitioners focus their efforts on preparing thoughtful and detailed educational sessions intended to increase program adoption, but do not allow time for interactive questions or develop opportunities for building relationships between key personnel responsible for successful adoption, then we see a reliance on the Transaction orientation to communication. When communication is conceptualized as a Process, we emphasize its constitutive nature wherein our environments – social, organizational, political, etc. – shape and are shaped through communication [ 3 ]. From a Process orientation, the transformative properties of communication emphasize its relational dimension and bring about a materiality from the intangible elements of the process (e.g., tone of voice, relational history, contextual exigency), and concepts such as psychological safety, mutual respect, and trust foreground the mechanics of information exchange. For example, someone may schedule multiple options for the same information session to ensure real-time interactivity for questions and build in opportunities for small group breakouts and post-presentation networking for relationship-building. When understanding of communication shifts to encompass more than information exchange, we begin to recognize the role of communication in building relationships and influencing long term cultural shifts, which is often the goal for implementation scientists [ 3 ]. If the Process orientation is overlooked in favor of a Transaction orientation, we may miss opportunities for identifying evidence-based communication strategies to support implementation.

The majority of subsequent work engaging Manojlovich et al.’s assertions agree on the imperative to engage a Process orientation to communication, but they make no strides in designing approaches for exploring the characteristics of communication surrounding effective implementation strategies (e.g., [ 6 , 7 , 8 ]). As the conversation initiated by Manojlovich and colleagues about the role of communication in implementation science has progressed, recognition of communication has grown, but emphasis continues to focus on formal contexts (e.g., trainings and webinars) [ 1 ]. Further, quantitative measures that assess information accuracy like the one used in Zhao and colleagues’ work overlook the importance of informal communication (e.g., rapport-building before meetings, impromptu connections) and the nuanced influence of the relational dimension that contributes to effective implementation. Bustos et al.’s (2021) analysis acknowledges both the formal and informal strategies through which communication might occur, but the communication they refer to is discussed from a Transaction orientation (i.e., “how information… was communicated to program staff” (p. 10)) [ 9 ].

For this study, we draw on interviews with employees of the Department of Veterans Affairs (VA) who evaluated, implemented, and administered interventions focused on improving the health and well-being of rural Veterans or the clinical staff who serve them. These interviews were exploratory and wide-ranging; for the purposes of this manuscript, we treat the interviews as akin to direct observations of intervention stakeholders discussing their real-world experiences operationalizing implementation strategies. Instead of focusing on what we could learn from the communication described in the interviews, we directed our attention to what lessons could be missing because of the way participants discussed communication. In this manuscript, we provide examples of how Transaction and Process orientations to communication appear in the data when individuals described their experiences, as well as their relationships that supported IS strategies and facilitated intervention goals. We also suggest interview strategies to elicit detail about communication from a Process orientation to support ongoing learning of these informal communication processes. Though these interviews were not focused on communication, we use data from the interviews to argue that noticing communication helps us discover how to do implementation science better. Specifically, a Process orientation emphasizes the space between IS strategies and outcomes, and advances understanding of implementation challenges and solutions.

Study setting and context

The VA’s Office of Rural Health (ORH) supports the creation of Enterprise-Wide Initiatives (EWIs) to address issues facing rural Veterans from mental health and primary care access to training and education of VA staff who serve rural Veterans. As a part of the funding cycle, EWI teams must conduct annual evaluations. The Center for the Evaluation of Enterprise-Wide Initiatives (CEEWI) was created through a 2019 partnership between ORH and the VA’s Quality Enhancement Research Initiative to support EWI evaluation and disseminate best practices. The CEEWI team, consisting of implementation science experts and qualitative data analysts, reviews the annual reports and provides feedback to EWI teams on reporting standards.

Data collection

As part of the initial CEEWI project, EWI evaluators, implementors, and administrators were interviewed about effectiveness of IS strategies they used and why, in part, to assist the CEEWI team in understanding key aspects of EWI implementation and evaluation. The interview guide included questions about the participant’s role on the EWI, the core components of the EWI, implementation strategies and their impact on desired outcomes, outcome measures used for evaluation, and the evaluation process. CEEWI team members and EWI leadership identified the evaluators, implementors, and administrators to recruit for the study. While recruitment sought a purposive sample of roles from each EWI, ultimately the sample was a convenience sample based on availability and willingness to participate during the first nine months of the COVID-19 pandemic. Additional details about recruitment and data collection can be found in an earlier manuscript from this larger project [ 10 ]. We conducted 43 semi-structured interviews, which averaged 51 min (range 20–77 min), from April – December 2020 with evaluators, implementors, and administrators from 21 EWIs. While most interviews were conducted one-on-one, 8 were group interviews ranging from 2 to 4 participants [ 10 ]. This study uses these interviews as an example on how communication is described when discussing implementation strategies.

Data analysis

Audio-recordings were transcribed, reviewed for accuracy, and uploaded into MAXQDA, a qualitative data management software [ 11 ]. Two doctorally trained qualitative analysts (NJ & JVT) leveraged their previous IS knowledge and conducted primary-cycle inductive coding to identify IS constructs and trends in the data [ 12 ]. The analysts initially coded all transcripts together in real-time and resolved discrepancies immediately. During this first round of coding, several IS constructs were identified in participants’ discussion of their implementation strategies, including Staff Buy-in, Tailoring, Rapport, Fidelity, and Mentorship. Collaboration and Leadership Buy-in emerged as the two most discussed IS constructs among participants. For secondary-cycle deductive coding to interpret how communication was conceptualized in discussions of Collaboration and Leadership Buy-in, the lead author, a Health Communication scholar, used an iterative process to develop a codebook to identify the language representing a Process or Transaction orientation for each construct (i.e., Collaboration and Leadership Buy-in) [ 3 , 12 ]. The analysis focused on the how communication was discussed, not about the form of communication that took place.

Collaboration, a term often characterizing various levels of formal and informal partnerships between individuals, departments or organizations, is defined as a mutually beneficial and well-defined relationship between two or more parties to achieve common goals [ 13 ]. An example of discussing Collaboration from a Transaction orientation to communication would be using the term Collaboration to describe monthly meetings where the parties update one another about the status of their tasks and goals. From a Process orientation, Collaboration would be discussed in relational terms, describing the trust and rapport the team members have among one another.

Leadership Buy-in represents the role of support from individuals in leadership positions for a program’s adoption and sustainability, particularly when competing clinical and administrative demands are at play [ 7 ]. An example of discussing Leadership Buy-in using a Transaction orientation to communication would be a description of strategies for adoption that only focused on leadership education. However, someone who engaged a Process orientation to communication might: (1) discuss tailored persuasive strategies for demonstrating value to specific decision-makers, or (2) acknowledge the necessity for long-term relationships with individuals in leadership roles for sustainment.

We conducted 43 interviews with 28 evaluators, 12 implementors, and 12 administrators. We coded a total of 90 segments as Collaboration ( n  = 46) and Leadership Buy-in ( n  = 44) across all the interviews. Most segments coded as Collaboration ( n  = 34, 74%) and Leadership Buy-in ( n  = 31, 70%) discussed communication from a Transaction orientation. The following results present examples of the discussion of Collaboration and Leadership Buy-in from the Transaction and Process orientations to communication.

Transaction orientation to communication

When communication is treated as a transaction, it is discussed as a one-way flow of information traveling from one party to another during a discrete moment in time [ 3 ]. The materiality of communication is reduced to accurate, efficient information transfer, thus putting emphasis on the channel (e.g., Teams meeting, email) through which information is exchanged and the task dimension of the interaction.

Collaboration as transaction

Participants sometimes discussed Collaboration in a way that missed its nuance and treated communication as merely a means for transferring information that produced Collaboration. For example, one participant implied that communication, regardless of quality, is inherently good, thus the more there is, the better. They identified “communication across the team level” as an important strategy having the most impact on desired outcomes. “The more communication there is, the more people are able (…) to divide up [responsibilities].” (1A) In this instance, communication is synonymous with information exchange. While we do not have enough information to assess the quality of communication that Participant 1A is referring to, the fact they only discussed the parties involved and quantity of communication is an example of the Transaction orientation to communication.

In another example, a participant explained what they felt did not work as well in their evaluation process. “We have excellent communication with some, but not all members of the [EWI] (…) I’m not sure they’re always on the same page with each other, and then depending on who we’re having a meeting with, we might hear one thing but then that’s not what someone else was going to do (…) that’s one of the pieces that I think is hard for us.” (2A) Again, we see the Transaction orientation, and the barometer for effective communication is accuracy. The participant went on to discuss ways to improve this lack of alignment among team members, suggesting that “even if it’s just being invited to join calls (…) [for us] to answer questions about the [evaluation] data” would improve teamwork. (2A) This passage highlights an important aspect of communication – being present for an interaction and having the opportunity to answer questions enables information exchange.

One participant described the communication that occurred during a monthly videoconference:

The learning collaborative is focused on bringing people [together] to share their experiences and how various facilitators identify ways to shape their program, but also the way that our national team gives feedback about the data (…) One call a month is right after a report (…) they do a data review on the call where they go over the numbers with the entire learning collaborative, everyone in the program, giving them feedback from a national perspective and always reminding people of the milestones of the metrics that they’ve agreed to under the ORH grant. (3A)

Here, we see another example of a participant discussing communication in terms of information exchange.

Leadership buy-in as transaction

Participants also discussed Leadership Buy-in from a Transaction orientation. In the following passage, participant 4A described the benefits of the EWI leadership team visiting sites in-person:

They would do a site visit to all the hubs (…) and meet with the local leadership team and that’s where they confirmed if there were any issues that they might have. They would do like a 2–3 day site visit (…) so it helped create that structure where people knew exactly who to report to and how these programs were established and plenty of opportunities to address any concerns or any issues they might have.

There are substantial implications for local Leadership Buy-in through in-person visits, yet the only aspect of communication discussed here is information exchange and clarifying the information flow hierarchy (i.e., who to report to).

Participant 5A described their program’s efforts to obtain Leadership Buy-in:

Simple outreach and education, that was really the only things that we could do, and then as they continued, training kind of showed its usefulness. That had an impact on leadership buy-in.

Here, buy-in is attributed to education, which may account for some or even most of buy-in, but it does not recognize the relational dimension of communication.

For another EWI, leadership turnover at the facility presented a significant barrier to program sustainment, because Leadership Buy-in was perpetually reset, which exacerbated a “conflict between implementation and sustainment strategies” when the decision-maker for sustainment funding was not the same person to “sign off on it originally” (9A). Given the EWI provided seed-funding for specialty staff to implement the program, the expectation was that the facility would eventually incur the expense for sustainment, but the plan for funds was not made explicit at the time of application for the seed-funding. Participant 9A went on to explain how their program responded to the unforeseen challenge obtaining sustainment funding from sites:

Our clinical director worked really hard with the first cohort of sites prior to their funding ending to try to come up with strategies to pitch the program to leadership (…) Most sites had challenges with changing leadership priorities.

In response, the interviewer clarified their sources for funding, then changed topics: “Interviewer: Ok, alright. How about strategies that were intended to optimize the effectiveness outcomes for your EWI?” In this example, the interviewer seems to be approaching the participant’s description of Leadership Buy-in from a Transactional orientation. A Process-oriented approach that asked about the nature and details of pitching the EWI to leadership may have provided more information about implementation strategy.

Process orientation to communication

From the process perspective, no single interaction serves as the cause or proof of effective Collaboration. Rather, the Process orientation recognizes the value of communication lies in the cumulative outcomes of consistent, often routine, interactions.

Collaboration as process

Collaborations require shared responsibility, mutual authority, accountability, and sharing of resources and rewards for success [ 13 ]. Collaboration in implementation has focused on strategies to enhance partners’ ability to work together to achieve mutual benefits. We identified examples from participants discussing Collaboration with a Process orientation to communication. From these examples we see that Collaboration is seen as a product of long-term efforts to develop positive relationships and establish trust and autonomy to make one’s own decisions. Many participants recognized the uniqueness and value in reaching the point of Collaboration. For example, Participant 10A shared, “The partnerships, it’s like a very special kind of relationship–, where we have to trust them, we rely on each other, but we also need to be able to make independent decisions.” Participant 6A also recognized the importance of relationships, “I would say they’re collegial but they’re not fully collaborative (…) when they’re really more deeply integrated and their role is understood and recognized (…) they are more collaborative members.”

One participant on a different EWI echoed this sentiment that individuals’ intent and motivations for the work should extend beyond the assignment to be considered Collaboration, “It’s not just trying to check off a box (…) there truly is a passion behind it, on all of our parts, and that has been wonderful.” (7B) Recognizing others’ intent for their work allows one to acknowledge how interpersonal communication is influenced by more than information exchange.

In the following exemplars, we can see how interviewers were able to elicit detail about the interactions surrounding the implementation strategies they were discussing.

Exemplar for Probing Collaboration . In Table  1 , we share an exemplar for engaging the Process orientation to communication, which led to greater explication of the role of communication in the implementation process.

Through this example, we see a more nuanced treatment of communication as a process after the interviewer probed twice to understand the participant’s use of “facilitation” as an implementation strategy. We gained description of the collaborative atmosphere within a team and how individuals’ psychological safety is manifested through authentic interactions.

Leadership buy-in as process

It takes more than information-exchange to garner support (e.g., financial, staff) for facilitation and sustainment. One participant acknowledged the web of influence that contributes to Leadership Buy-in and effective implementation:

We reached out to all the rural sites their leadership… sort of advertising the program, so we would schedule a conference call with a director, chief of staff, emergency room chief, to sort of discuss the program (…) then we would follow up with an actual 1-day on-site visit (…) where we meet with again, leadership, but we also meet with the [staff from several departments] (…) It’s an all-day visit to further introduce our program, to the team on site, as well as learn more about their program, and how [our EWI] might incorporate itself, and what challenges (…) we might face in implementation. (2B)

Here, we see an acknowledgement of reciprocal relationship-building to learn about priorities and needs.

Several participants discussed how time costly it is to gain Leadership Buy-in to ease the burden of change on an organization and staff, particularly for a nationwide program. One participant reflected:

Ten years ago, it was a [regional] project, so the main kind of instruction came from a [regional] level down, you know. The site visit was just a medical director and the nurse manager telling you that, ‘Hey, this is what’s going to happen,’ and it happened. Now (…) it’s like a year-long process to get people familiarized (…) go live went from one day to four days long. (11B)

Despite its value, garnering Leadership Buy-in has its challenges. Sometimes identifying the right individuals who represent the relevant leadership roles is not clear cut.

Once we have identified that our program can go to that site, we ask the local (…) program manager to identify who (…) key local leaders are (…) It’s important to have the managers of those sites involved in this process from the beginning (…) We (…) set up an initial meeting (…) where we review the implementation process plan with everybody on that call, and answer questions about what we and [specialty care] services will provide as part of the training opportunity and clearly delineate what we need the site or the facility to commit to provide (…) we answer questions, alleviate concerns, things like that. (7B)

Participant 7B went on to describe the challenge of identifying the right leadership representative:

The only barrier that we’ve encountered is some challenges in getting the right leadership on the call to review this in real time and answer questions (…) whether it is due to leadership turnover at the site, even from the time that we set up the call to the time that we actually do the call, there have been some change-overs, and that has been a challenge.

Again, we see this participant engaging a strong Process orientation to communication as they emphasize the importance of relationship-building for Leadership Buy-in.

Exemplar for Probing Leadership Buy-in . In the following example, the interviewer engaged the Process orientation to communication with probes that led to greater explication of the role of communication in developing Leadership Buy-in (Table  2 ).

Results illustrate ways administrators, implementors, and evaluators characterized communication related to Collaboration and Leadership Buy-in. From the Transaction orientation, we saw that the term communication was used synonymously for information exchange. The problem of implementation lies beyond efficient and reliable information transfer, and instead centers on cooperative sensemaking and learning within and among teams situated in an organization that is influenced by its social, geographic, and political environments [ 2 , 14 , 15 ]. Communication necessary for effective implementation is based on improvisation and reciprocity and constitute relationships over time [ 2 , 15 ]. Our data indicate these processes are occurring in implementation, but we may not always be paying close enough attention to their occurrence. If most discussions about communication engage a Transaction orientation, then practitioners and evaluators will never have the insight necessary to maximize the impact of their communication efforts.

Participants often discussed Leadership Buy-in more as an outcome of education, and less as a byproduct of improvisational relationship-building, which demonstrates the predominant Transaction orientation to communication privileging rehearsed, often unidirectional, and mostly controlled interactions. Formal information exchange is undoubtedly an important element of effective implementation; the Transaction orientation aligns well with the goals of dissemination and implementation as a field [ 15 ]. However, our data point to the importance of thinking about communication from a Process orientation for improving effectiveness of implementation strategies—and show how members of implementation and evaluation teams too often focus on the transaction elements of communication. Previous work that engages the Transaction orientation and points to the benefits of reliable information exchange has paved the way for more exploratory naturalistic methods for studying IS from a Process orientation to communication [ 3 , 14 , 15 , 16 ]. As noted in our findings, the Transaction orientation overlooks the intricacies of processes that occur among individuals to build trust, cultivate buy-in, and influence team decision-making, all of which are markers of successful implementation.

Suggestions for engaging process orientation to communication

Given the purpose of IS is to promote the adoption of research and evidence-based practices, it would behoove implementation scientists to tap into the richness of interdisciplinary theorizing and engage a Process orientation to communication [ 17 ]. As thinking about communication has evolved from a Transaction orientation, scholars recognized the symbolic process that humans use to create meaning through informal, improvised interactions over a period of time [ 2 ]. Recent analysis of implementation strategies for behavioral health interventions called for explicit attention to the supportive role communication may play in most, if not all, strategies [ 15 ]. The Process orientation to communication enriches theorizing and elevates scholars’ and practitioners’ understanding of how to leverage implementation strategies to be meaningfully responsive to the relationships among the interested parties [ 18 ]. However, we warn against over-characterizing communication into a ‘nebulous, global process’ [ 2 , 19 ]. For gaining insight on communication processes, we suggest two strategies: 1) interviewers focused on understanding implementation strategies could probe their interviewees to learn more about how communication is enacted; and 2) IS practitioners could utilize process-oriented communication theories in developing interventions and evaluation tools (e.g., interview guides).

The supplementary material accompanying this article includes excerpts from our interview data as examples demonstrating hypothetical ways interviewers can elicit more nuanced understanding of communication processes (see Tables S1 and S2).

Our analysis identified examples of missed opportunities for interviewers to probe about communication from a Process orientation recognizing the relational dimension of communication. Interview probes like those recommended in Tables S1 and S2 could lead to valuable understanding of the processes of communication, allowing exploration of the relational dimension of communication and implementation, and insight to individuals’ attitudes and sensemaking about those experiences. This may contribute to a more nuanced understanding of the importance of communication in implementation strategies beyond a transactional information exchange. We also provided examples highlighting the constitutive role communication plays in relationship-building. Our goal is to help attune IS researchers to the value of the processes of communication as a critical component of many implementation strategies.

Probing for communication processes in interviews

Challenges to implementing any new program may be significantly varied and widespread. No single barrier serves as an intervention’s fatal flaw, but rather, implementation is affected by numerous factors shaped through informal interactions [ 17 , 20 ]. A recent study that aimed to identify which implementation strategies should be most closely considered for which determinants of practice reported one of its limitations was the heterogeneity of responses [ 21 ]. This variation in responses among administrators, implementors, and evaluators points to the value of a more nuanced understanding of the unique, context-dependent, and relationally based communication processes undergirding implementation strategies [ 21 ]. Further, in their ethnographic study on hand hygiene programs, Goedken and colleagues poignantly emphasized the importance of understanding how implementation strategies are used and defined in real-world settings for understanding determinants of practice [ 22 ]. By looking below the surface of implementation strategies and focusing on the interactions surrounding those strategies, we may begin to recognize the determinants of practices, the mechanisms for change, more precisely. Discussing communication from a Process orientation allows us to access what is happening below the surface that cannot be observed as an outsider. With greater insight on communication processes occurring throughout implementation, the field of IS would be poised to provide meaningful guidance for combining implementation strategies [ 22 ]. In a similar vein, IS researchers should consider the temporality of IS strategies and how this underscores the role of communication. The role of Leadership Buy-in at all stages of development and implementation on effectiveness cannot be overstated [ 23 ]. Albright suggests shifting away from the predominant focus of research on the active implementation period to explore activities occurring during design and preparation [ 15 ].

Most implementation strategies have a communication component representing the channel for education and promotion (e.g., workshops, webinars, brochures) [ 15 ]. Our proposed interview strategies interrogate communication in a way that recognizes the relational dimensions of interpersonal interactions, providing insight about what truly results in effective implementation. By understanding communication from a Process orientation, we may enrich our understanding of implementation strategies [ 24 ].

Utilize process-oriented theories

Theories that engage a Transaction orientation to communication often ascribe to the traditional knowledge-intention-behavior paradigm that proposes a stable, linear positive relationship between knowledge and behavior change (e.g., Theory of Reasoned Action, a predictive theory suggesting a strong relationship among individuals’ attitudes about a behavior, their intention, and their behavior [ 25 ]) and tends to overlook the nuance of communication processes. However, humans are more complicated and inconsistent than these theories acknowledge. The Process orientation to communication allows for more realistic approaches that privilege the constitutive nature of communication to co-create meaning socially. In a recent scoping review of 158 studies in implementation research on maternity care, effective communication was noted as a key factor for promoting change across the body of work, but the majority of research was atheoretical and ambiguous in operationalization of communication [ 26 ].

Health communication scholars are trained to be sensitive to the cooperative nature of establishing shared meaning, multiple interpretations of behaviors, and the challenges of coordinating interactions when studying implementation strategies. Several theories, including two that pay special attention to how meaning is created socially, Coordinated Management of Meaning (CMM) [ 27 ] and Structuration Theory [ 28 ], could highlight perspectives that recognize communication as a complex process and translate well to practice. CMM is a constructivist theory that provides a practical heuristic for interpreting interpersonal communication events that comprise larger conversations. As such, CMM informs practitioners’ decision-making by illuminating patterns of interactions to find ways of talking that could result in desired outcomes [ 29 ]. Structuration Theory, coined by sociologist Anthony Giddens in the late 1970s, describes the dynamic relationship between individuals and their environment that constrains and enables social practices [ 28 ]. Through its critical lens, Structuration Theory highlights the (lack of) agency individuals perceive for themselves and others, and the rules and resources perpetuated through social interactions. Lastly, Diffusion of Innovations, a framework well-entrenched in IS research and practice, also engages a process paradigm [ 30 , 31 ]. There is ample opportunity and an imperative to employ a Process orientation to better understand communication in implementation science.

Limitations

This study has multiple limitations. We did not collection demographic data to describe our participants beyond the role they held on their EWI teams. The data represents a convenience sample of administrators, implementors, and evaluators working on EWIs funded at the time of data collection, which resulted in variability in representation across EWIs and staff roles. Further, because of the diversity of foci, designs, and timelines of EWIs, we cannot draw conclusions about effectiveness of strategies discussed in this paper. Lastly, the interviews were not conducted to assess communication explicitly. Despite these limitations, our analysis facilitates concrete suggestions for improving understanding of the role of communication in implementation.

Future directions for research

Research analyzing the role of communication from a Process orientation would enrich the field of IS. Similar to Fishman et al.’s work comparing measurement and operationalization of attitude among IS studies and those grounded in psychology, our work emphasizes the importance of interdisciplinary collaboration [ 32 ]. The interviewees and interviewers in our study focused predominantly on a Transaction orientation to communication; more studies are needed that focus on this level of distinction, particularly how to adopt a Process orientation to communication for implementation strategy specification. There is great potential for a body of knowledge about communication processes that has been systematically developed to inform IS strategies supporting a range of aspects crucial to effectiveness including Leadership Buy-in and Collaboration. Future research may do well to conduct direct observation to characterize communication processes related to implementation strategies from a rich Process orientation. Dissemination Science, as one facet of Dissemination and Implementation Science, is firmly rooted in the mechanics of communication and would greatly benefit from engaging the Process orientation. A recent scoping review demonstrated that the field of Dissemination Science lacks insight to communication from the Process orientation; in their review of dissemination determinants, the Transaction orientation persists in focusing on imparting information from one party to the next [ 33 ].

This study described instances of two broadly accepted orientations to communication engaged by implementation scientists. The findings demonstrate opportunities – and strategies – for engaging in the Process orientation of communication to gain greater insight into the role communication plays in implementation outcomes. We hope this work inspires dialogue, new interdisciplinary collaboration, and innovative methods to highlight the utility of engaging the Process orientation to communication to undergird the value of communication theory to implementation science for improving health services. When communication is understood as a process, practitioners will be better able to prepare for the unpredictability and uniqueness of the relational dimensions of communication.

Availability of data and materials

The datasets presented in this article are not readily available in accordance with federal requirements and standards and guidelines for the protection of participants’ privacy and to maintain confidentiality. Requests to access the datasets should be directed to Dr. Heather Reisinger ([email protected]).

Abbreviations

Center for the Evaluation of Enterprise-Wide Initiatives

Coordinated Management of Meaning

Enterprise-Wide Initiative

Implementation Science

Office of Rural Health

Department of Veterans Affairs

Zhao X, et al. Perceived communication effectiveness in implementation strategies: a measurement scale. Implement Sci Commun. 2022;3(1):38.

Article   PubMed   PubMed Central   Google Scholar  

Jordan ME, et al. The role of conversation in health care interventions: enabling sensemaking and learning. Implement Sci. 2009;4: 15.

Manojlovich M, et al. Hiding in plain sight: communication theory in implementation science. Implement Sci. 2015;10:58.

What is communication? 2023. Available from: https://www.natcom.org/about-nca/what-communication . Cited 2023 July 11.

Kreps G. Analysis of the interdisciplinary credibility of communication as a social science. Assoc Communication Bull. 1982;42:40–3.

Google Scholar  

Lette M, et al. Unknown makes unloved-a case study on improving integrated health and social care in the Netherlands using a participatory approach. Health Soc Care Community. 2020;28(2):670–80.

Article   PubMed   Google Scholar  

Lewis MA, et al. Multilevel communication to improve well-being during a pandemic. Implement Res Pract. 2021;2:2633489520988265.

Salvador JG, et al. Use of concept mapping to support evidence-based practice implementation improvement in rural areas. J Rural Mental Health. 2018;42(1):3–19.

Article   Google Scholar  

Bustos TE, Sridhar A, Drahota A. Community-based implementation strategy use and satisfaction: a mixed-methods approach to using the ERIC compilation for organizations serving children on the autism spectrum. Implement Res Pract. 2021;2:26334895211058086.

Chasco EE, et al. RE-AIM for rural health innovations: perceptions of (mis) alignment between the RE-AIM framework and evaluation reporting the Department of Veterans Affairs enterprise-wide Initiative program. Front Health Serv. 2024;4:1278209.

MAXQDA Plus 2022. Version 22.3.0. VERBI GmbH 1995–2023. Updated 2023. https://www.maxqda.com . Accessed 4 Mar 2023.

Saldana J. The coding manual for qualitative researchers. Sage; 2009.

Huang KY, et al. Unpacking partnership, engagement, and collaboration research to inform implementation strategies development: theoretical frameworks and emerging methodologies. Front Public Health. 2018;6: 190.

Kislov R, et al. Harnessing the power of theorising in implementation science. Implement Sci. 2019;14(1):103.

Albright K, et al. Communication strategies to facilitate the implementation of new clinical practices: a qualitative study of community mental health therapists. Transl Behav Med. 2022;12(2):324–34.

Salas E, et al. Communicating, coordinating, and cooperating when lives depend on it: tips for teamwork. Jt Comm J Qual Patient Saf. 2008;34(6):333–41.

PubMed   Google Scholar  

Eccles MP, et al. An implementation research agenda. Implement Sci. 2009;4: 18.

Ornstein JT, et al. Rugged landscapes: complexity and implementation science. Implement Sci. 2020;15(1):85.

Donovan EE, LeBlanc K, Farris. Interpersonal communication and coping with cancer: a multidisciplinary theoretical review of the literature. Communication Theory. 2019;29(2):236–56.

McCullough MB, et al. The interplay of contextual elements in implementation: an ethnographic case study. BMC Health Serv Res. 2015;15:62.

Waltz TJ, et al. Choosing implementation strategies to address contextual barriers: diversity in recommendations and future directions. Implement Sci. 2019;14(1):42.

Goedken CC, et al. The role as a champion is to not only monitor but to speak out and to educate: the contradictory roles of hand hygiene champions. Implement Sci. 2019;14(1):110.

Ross J, et al. Factors that influence the implementation of e-health: a systematic review of systematic reviews (an update). Implement Sci. 2016;11(1):146.

Baker R, et al. Tailored interventions to address determinants of practice. Cochrane Database Syst Rev. 2015;2015(4):CD005470.

PubMed   PubMed Central   Google Scholar  

Montano DE, Kasprzyk D. Theory of reasoned action, theory of planned behavior, and the integrated behavioral model. In: Glanz K, Rimer BK, Viswanath K, editors. Health behavior: theory, research, and practices. Jossey-Bass; 2015. p. 75–94.

Dadich A, Piper A, Coates D. Implementation science in maternity care: a scoping review. Implement Sci. 2021;16(1):16.

Pearce WB, Pearce KA. Extending the theory of the coordinated management of meaning (CMM) through the community dialogue process. Communication Theory. 2000;10(4):405–23.

Hardcastle MA, Usher KJ, Holmes CA. An overview of structuration theory and its usefulness for nursing research. Nurs Philos. 2005;6(4):223–34.

Imran M, et al. A critical study of coordinated management of meaning theory: a theory in Practitioners’ hands. Int J Engl Linguistics. 2019;9(5):301.

Rogers EM. Diffusion of innovations. 5th ed. Free Press; 2003.

Powell BJ, et al. A mixed methods multiple case study of implementation as usual in children’s social service organizations: study protocol. Implement Sci. 2013;8(1):92.

Fishman J, Yang C, Mandell D. Attitude theory and measurement in implementation science: a secondary review of empirical studies and opportunities for advancement. Implement Sci. 2021;16(1):87.

Baumann AA, et al. A scoping review of frameworks in empirical studies and a review of dissemination frameworks. Implement Sci. 2022;17(1):53.

Download references

Acknowledgements

We would like to thank the interview participants who participated in this study for their time and insights. We would also like to acknowledge Office of Rural Health (ORH) program analysts Dr. Kelly Lora Lewis, Karyn Johnstone, Nicole Sanchez, Maura Timm, Anthony Achampong, Richard Huang, and Janice Garland for their assistance, as well as Dr. Sheila Robinson, former Deputy Director of ORH, Dr. Peter Kaboli, Executive Director of ORH, and Dr. Thomas Klobucar, former Executive Director of ORH, for their support. The views expressed in this article are those of the authors and do not necessarily reflect the position or policy of the Department of Veterans Affairs or the United States government.

VA Office of Rural Health and QUERI Project #: PEC 19–456.

Author information

Authors and affiliations.

Center for Access and Delivery Research and Evaluation, Iowa City VA Healthcare System, Iowa City, IA, USA

Nicole L. Johnson, Jennifer Van Tiem, Erin Balkenende, DeShauna Jones, Julia E. Friberg, Emily E. Chasco, Jane Moeckli, Kenda S. Steffensmeier, Melissa J. A. Steffen & Heather Schacht Reisinger

Veterans Rural Health Resource Center-Iowa City (VRHRC-Iowa City), VA Office of Rural Health, Iowa City, IA, USA

Nicole L. Johnson, Jennifer Van Tiem, Julia E. Friberg, Jane Moeckli, Kenda S. Steffensmeier & Melissa J. A. Steffen

Division of General Internal Medicine, Department of Internal Medicine, Carver College of Medicine, University of Iowa, Iowa City, IA, USA

Erin Balkenende & Heather Schacht Reisinger

Institute for Clinical and Translational Science, University of Iowa, Iowa City, IA, USA

DeShauna Jones, Emily E. Chasco & Heather Schacht Reisinger

Department of Health Management and Policy, College of Public Health, University of Iowa, Iowa City, IA, USA

Kanika Arora

Herbert Wertheim School of Public Health and Human Longevity Science, University of California San Diego, La Jolla, CA, San Diego, USA

Borsika A. Rabin

UC San Diego ACTRI Dissemination and Implementation Science Center, University of California San Diego, La Jolla, CA, San Diego, USA

You can also search for this author in PubMed   Google Scholar

Contributions

HSR, EC, JVT, NJ, EB, DJ, and JF are responsible for the concept for this manuscript. NJ drafted the initial manuscript and HSR, JVT, EC, EB, DJ, KSS, and JF contributed substantially in the form of manuscript structure and revisions. HSR developed the proposal for this project and obtained funding, with input from JVT, EB, and JM. HSR, JVT, EB, JM, and MS conducted interviews. KA and BR advised on all aspects of the project including development of the standardized evaluation reporting template and manuscript revisions.

Corresponding author

Correspondence to Nicole L. Johnson .

Ethics declarations

Ethics approval and consent to participate.

As designated by the IRB, this review is not human subject research (Protocol #202001043).

Consent for publication

Not applicable.

Competing interests

HSR is an associate editor of Implementation Science . All decisions on this paper were made by another editor. The authors declare that they have no other competing interests.

Additional information

Publisher’s note.

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

Supplementary Information

Supplementary material 1. , 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/ .

Reprints and permissions

About this article

Cite this article.

Johnson, N.L., Van Tiem, J., Balkenende, E. et al. Gaps in communication theory paradigms when conducting implementation science research: qualitative observations from interviews with administrators, implementors, and evaluators of rural health programs. Implementation Sci 19 , 66 (2024). https://doi.org/10.1186/s13012-024-01395-3

Download citation

Received : 07 May 2024

Accepted : 03 September 2024

Published : 16 September 2024

DOI : https://doi.org/10.1186/s13012-024-01395-3

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Leadership buy-in
  • Collaboration
  • Interviewing
  • Qualitative methods
  • Communication theory
  • Implementation science strategies

ISSN: 1748-5908

  • Submission enquiries: Access here and click Contact Us
  • General enquiries: [email protected]

qualitative meaning of research

What is Qualitative in Qualitative Research

  • Open access
  • Published: 27 February 2019
  • Volume 42 , pages 139–160, ( 2019 )

Cite this article

You have full access to this open access article

qualitative meaning of research

  • Patrik Aspers 1 , 2 &
  • Ugo Corte 3  

630k Accesses

354 Citations

25 Altmetric

Explore all metrics

What is qualitative research? If we look for a precise definition of qualitative research, and specifically for one that addresses its distinctive feature of being “qualitative,” the literature is meager. In this article we systematically search, identify and analyze a sample of 89 sources using or attempting to define the term “qualitative.” Then, drawing on ideas we find scattered across existing work, and based on Becker’s classic study of marijuana consumption, we formulate and illustrate a definition that tries to capture its core elements. We define qualitative research as an iterative process in which improved understanding to the scientific community is achieved by making new significant distinctions resulting from getting closer to the phenomenon studied. This formulation is developed as a tool to help improve research designs while stressing that a qualitative dimension is present in quantitative work as well. Additionally, it can facilitate teaching, communication between researchers, diminish the gap between qualitative and quantitative researchers, help to address critiques of qualitative methods, and be used as a standard of evaluation of qualitative research.

Similar content being viewed by others

qualitative meaning of research

What is Qualitative in Research

Unsettling definitions of qualitative research, what is “qualitative” in qualitative research why the answer does not matter but the question is important, explore related subjects.

  • Artificial Intelligence

Avoid common mistakes on your manuscript.

If we assume that there is something called qualitative research, what exactly is this qualitative feature? And how could we evaluate qualitative research as good or not? Is it fundamentally different from quantitative research? In practice, most active qualitative researchers working with empirical material intuitively know what is involved in doing qualitative research, yet perhaps surprisingly, a clear definition addressing its key feature is still missing.

To address the question of what is qualitative we turn to the accounts of “qualitative research” in textbooks and also in empirical work. In his classic, explorative, interview study of deviance Howard Becker ( 1963 ) asks ‘How does one become a marijuana user?’ In contrast to pre-dispositional and psychological-individualistic theories of deviant behavior, Becker’s inherently social explanation contends that becoming a user of this substance is the result of a three-phase sequential learning process. First, potential users need to learn how to smoke it properly to produce the “correct” effects. If not, they are likely to stop experimenting with it. Second, they need to discover the effects associated with it; in other words, to get “high,” individuals not only have to experience what the drug does, but also to become aware that those sensations are related to using it. Third, they require learning to savor the feelings related to its consumption – to develop an acquired taste. Becker, who played music himself, gets close to the phenomenon by observing, taking part, and by talking to people consuming the drug: “half of the fifty interviews were conducted with musicians, the other half covered a wide range of people, including laborers, machinists, and people in the professions” (Becker 1963 :56).

Another central aspect derived through the common-to-all-research interplay between induction and deduction (Becker 2017 ), is that during the course of his research Becker adds scientifically meaningful new distinctions in the form of three phases—distinctions, or findings if you will, that strongly affect the course of his research: its focus, the material that he collects, and which eventually impact his findings. Each phase typically unfolds through social interaction, and often with input from experienced users in “a sequence of social experiences during which the person acquires a conception of the meaning of the behavior, and perceptions and judgments of objects and situations, all of which make the activity possible and desirable” (Becker 1963 :235). In this study the increased understanding of smoking dope is a result of a combination of the meaning of the actors, and the conceptual distinctions that Becker introduces based on the views expressed by his respondents. Understanding is the result of research and is due to an iterative process in which data, concepts and evidence are connected with one another (Becker 2017 ).

Indeed, there are many definitions of qualitative research, but if we look for a definition that addresses its distinctive feature of being “qualitative,” the literature across the broad field of social science is meager. The main reason behind this article lies in the paradox, which, to put it bluntly, is that researchers act as if they know what it is, but they cannot formulate a coherent definition. Sociologists and others will of course continue to conduct good studies that show the relevance and value of qualitative research addressing scientific and practical problems in society. However, our paper is grounded in the idea that providing a clear definition will help us improve the work that we do. Among researchers who practice qualitative research there is clearly much knowledge. We suggest that a definition makes this knowledge more explicit. If the first rationale for writing this paper refers to the “internal” aim of improving qualitative research, the second refers to the increased “external” pressure that especially many qualitative researchers feel; pressure that comes both from society as well as from other scientific approaches. There is a strong core in qualitative research, and leading researchers tend to agree on what it is and how it is done. Our critique is not directed at the practice of qualitative research, but we do claim that the type of systematic work we do has not yet been done, and that it is useful to improve the field and its status in relation to quantitative research.

The literature on the “internal” aim of improving, or at least clarifying qualitative research is large, and we do not claim to be the first to notice the vagueness of the term “qualitative” (Strauss and Corbin 1998 ). Also, others have noted that there is no single definition of it (Long and Godfrey 2004 :182), that there are many different views on qualitative research (Denzin and Lincoln 2003 :11; Jovanović 2011 :3), and that more generally, we need to define its meaning (Best 2004 :54). Strauss and Corbin ( 1998 ), for example, as well as Nelson et al. (1992:2 cited in Denzin and Lincoln 2003 :11), and Flick ( 2007 :ix–x), have recognized that the term is problematic: “Actually, the term ‘qualitative research’ is confusing because it can mean different things to different people” (Strauss and Corbin 1998 :10–11). Hammersley has discussed the possibility of addressing the problem, but states that “the task of providing an account of the distinctive features of qualitative research is far from straightforward” ( 2013 :2). This confusion, as he has recently further argued (Hammersley 2018 ), is also salient in relation to ethnography where different philosophical and methodological approaches lead to a lack of agreement about what it means.

Others (e.g. Hammersley 2018 ; Fine and Hancock 2017 ) have also identified the treat to qualitative research that comes from external forces, seen from the point of view of “qualitative research.” This threat can be further divided into that which comes from inside academia, such as the critique voiced by “quantitative research” and outside of academia, including, for example, New Public Management. Hammersley ( 2018 ), zooming in on one type of qualitative research, ethnography, has argued that it is under treat. Similarly to Fine ( 2003 ), and before him Gans ( 1999 ), he writes that ethnography’ has acquired a range of meanings, and comes in many different versions, these often reflecting sharply divergent epistemological orientations. And already more than twenty years ago while reviewing Denzin and Lincoln’ s Handbook of Qualitative Methods Fine argued:

While this increasing centrality [of qualitative research] might lead one to believe that consensual standards have developed, this belief would be misleading. As the methodology becomes more widely accepted, querulous challengers have raised fundamental questions that collectively have undercut the traditional models of how qualitative research is to be fashioned and presented (1995:417).

According to Hammersley, there are today “serious treats to the practice of ethnographic work, on almost any definition” ( 2018 :1). He lists five external treats: (1) that social research must be accountable and able to show its impact on society; (2) the current emphasis on “big data” and the emphasis on quantitative data and evidence; (3) the labor market pressure in academia that leaves less time for fieldwork (see also Fine and Hancock 2017 ); (4) problems of access to fields; and (5) the increased ethical scrutiny of projects, to which ethnography is particularly exposed. Hammersley discusses some more or less insufficient existing definitions of ethnography.

The current situation, as Hammersley and others note—and in relation not only to ethnography but also qualitative research in general, and as our empirical study shows—is not just unsatisfactory, it may even be harmful for the entire field of qualitative research, and does not help social science at large. We suggest that the lack of clarity of qualitative research is a real problem that must be addressed.

Towards a Definition of Qualitative Research

Seen in an historical light, what is today called qualitative, or sometimes ethnographic, interpretative research – or a number of other terms – has more or less always existed. At the time the founders of sociology – Simmel, Weber, Durkheim and, before them, Marx – were writing, and during the era of the Methodenstreit (“dispute about methods”) in which the German historical school emphasized scientific methods (cf. Swedberg 1990 ), we can at least speak of qualitative forerunners.

Perhaps the most extended discussion of what later became known as qualitative methods in a classic work is Bronisław Malinowski’s ( 1922 ) Argonauts in the Western Pacific , although even this study does not explicitly address the meaning of “qualitative.” In Weber’s ([1921–-22] 1978) work we find a tension between scientific explanations that are based on observation and quantification and interpretative research (see also Lazarsfeld and Barton 1982 ).

If we look through major sociology journals like the American Sociological Review , American Journal of Sociology , or Social Forces we will not find the term qualitative sociology before the 1970s. And certainly before then much of what we consider qualitative classics in sociology, like Becker’ study ( 1963 ), had already been produced. Indeed, the Chicago School often combined qualitative and quantitative data within the same study (Fine 1995 ). Our point being that before a disciplinary self-awareness the term quantitative preceded qualitative, and the articulation of the former was a political move to claim scientific status (Denzin and Lincoln 2005 ). In the US the World War II seem to have sparked a critique of sociological work, including “qualitative work,” that did not follow the scientific canon (Rawls 2018 ), which was underpinned by a scientifically oriented and value free philosophy of science. As a result the attempts and practice of integrating qualitative and quantitative sociology at Chicago lost ground to sociology that was more oriented to surveys and quantitative work at Columbia under Merton-Lazarsfeld. The quantitative tradition was also able to present textbooks (Lundberg 1951 ) that facilitated the use this approach and its “methods.” The practices of the qualitative tradition, by and large, remained tacit or was part of the mentoring transferred from the renowned masters to their students.

This glimpse into history leads us back to the lack of a coherent account condensed in a definition of qualitative research. Many of the attempts to define the term do not meet the requirements of a proper definition: A definition should be clear, avoid tautology, demarcate its domain in relation to the environment, and ideally only use words in its definiens that themselves are not in need of definition (Hempel 1966 ). A definition can enhance precision and thus clarity by identifying the core of the phenomenon. Preferably, a definition should be short. The typical definition we have found, however, is an ostensive definition, which indicates what qualitative research is about without informing us about what it actually is :

Qualitative research is multimethod in focus, involving an interpretative, naturalistic approach to its subject matter. This means that qualitative researchers study things in their natural settings, attempting to make sense of, or interpret, phenomena in terms of the meanings people bring to them. Qualitative research involves the studied use and collection of a variety of empirical materials – case study, personal experience, introspective, life story, interview, observational, historical, interactional, and visual texts – that describe routine and problematic moments and meanings in individuals’ lives. (Denzin and Lincoln 2005 :2)

Flick claims that the label “qualitative research” is indeed used as an umbrella for a number of approaches ( 2007 :2–4; 2002 :6), and it is not difficult to identify research fitting this designation. Moreover, whatever it is, it has grown dramatically over the past five decades. In addition, courses have been developed, methods have flourished, arguments about its future have been advanced (for example, Denzin and Lincoln 1994) and criticized (for example, Snow and Morrill 1995 ), and dedicated journals and books have mushroomed. Most social scientists have a clear idea of research and how it differs from journalism, politics and other activities. But the question of what is qualitative in qualitative research is either eluded or eschewed.

We maintain that this lacuna hinders systematic knowledge production based on qualitative research. Paul Lazarsfeld noted the lack of “codification” as early as 1955 when he reviewed 100 qualitative studies in order to offer a codification of the practices (Lazarsfeld and Barton 1982 :239). Since then many texts on “qualitative research” and its methods have been published, including recent attempts (Goertz and Mahoney 2012 ) similar to Lazarsfeld’s. These studies have tried to extract what is qualitative by looking at the large number of empirical “qualitative” studies. Our novel strategy complements these endeavors by taking another approach and looking at the attempts to codify these practices in the form of a definition, as well as to a minor extent take Becker’s study as an exemplar of what qualitative researchers actually do, and what the characteristic of being ‘qualitative’ denotes and implies. We claim that qualitative researchers, if there is such a thing as “qualitative research,” should be able to codify their practices in a condensed, yet general way expressed in language.

Lingering problems of “generalizability” and “how many cases do I need” (Small 2009 ) are blocking advancement – in this line of work qualitative approaches are said to differ considerably from quantitative ones, while some of the former unsuccessfully mimic principles related to the latter (Small 2009 ). Additionally, quantitative researchers sometimes unfairly criticize the first based on their own quality criteria. Scholars like Goertz and Mahoney ( 2012 ) have successfully focused on the different norms and practices beyond what they argue are essentially two different cultures: those working with either qualitative or quantitative methods. Instead, similarly to Becker ( 2017 ) who has recently questioned the usefulness of the distinction between qualitative and quantitative research, we focus on similarities.

The current situation also impedes both students and researchers in focusing their studies and understanding each other’s work (Lazarsfeld and Barton 1982 :239). A third consequence is providing an opening for critiques by scholars operating within different traditions (Valsiner 2000 :101). A fourth issue is that the “implicit use of methods in qualitative research makes the field far less standardized than the quantitative paradigm” (Goertz and Mahoney 2012 :9). Relatedly, the National Science Foundation in the US organized two workshops in 2004 and 2005 to address the scientific foundations of qualitative research involving strategies to improve it and to develop standards of evaluation in qualitative research. However, a specific focus on its distinguishing feature of being “qualitative” while being implicitly acknowledged, was discussed only briefly (for example, Best 2004 ).

In 2014 a theme issue was published in this journal on “Methods, Materials, and Meanings: Designing Cultural Analysis,” discussing central issues in (cultural) qualitative research (Berezin 2014 ; Biernacki 2014 ; Glaeser 2014 ; Lamont and Swidler 2014 ; Spillman 2014). We agree with many of the arguments put forward, such as the risk of methodological tribalism, and that we should not waste energy on debating methods separated from research questions. Nonetheless, a clarification of the relation to what is called “quantitative research” is of outmost importance to avoid misunderstandings and misguided debates between “qualitative” and “quantitative” researchers. Our strategy means that researchers, “qualitative” or “quantitative” they may be, in their actual practice may combine qualitative work and quantitative work.

In this article we accomplish three tasks. First, we systematically survey the literature for meanings of qualitative research by looking at how researchers have defined it. Drawing upon existing knowledge we find that the different meanings and ideas of qualitative research are not yet coherently integrated into one satisfactory definition. Next, we advance our contribution by offering a definition of qualitative research and illustrate its meaning and use partially by expanding on the brief example introduced earlier related to Becker’s work ( 1963 ). We offer a systematic analysis of central themes of what researchers consider to be the core of “qualitative,” regardless of style of work. These themes – which we summarize in terms of four keywords: distinction, process, closeness, improved understanding – constitute part of our literature review, in which each one appears, sometimes with others, but never all in the same definition. They serve as the foundation of our contribution. Our categories are overlapping. Their use is primarily to organize the large amount of definitions we have identified and analyzed, and not necessarily to draw a clear distinction between them. Finally, we continue the elaboration discussed above on the advantages of a clear definition of qualitative research.

In a hermeneutic fashion we propose that there is something meaningful that deserves to be labelled “qualitative research” (Gadamer 1990 ). To approach the question “What is qualitative in qualitative research?” we have surveyed the literature. In conducting our survey we first traced the word’s etymology in dictionaries, encyclopedias, handbooks of the social sciences and of methods and textbooks, mainly in English, which is common to methodology courses. It should be noted that we have zoomed in on sociology and its literature. This discipline has been the site of the largest debate and development of methods that can be called “qualitative,” which suggests that this field should be examined in great detail.

In an ideal situation we should expect that one good definition, or at least some common ideas, would have emerged over the years. This common core of qualitative research should be so accepted that it would appear in at least some textbooks. Since this is not what we found, we decided to pursue an inductive approach to capture maximal variation in the field of qualitative research; we searched in a selection of handbooks, textbooks, book chapters, and books, to which we added the analysis of journal articles. Our sample comprises a total of 89 references.

In practice we focused on the discipline that has had a clear discussion of methods, namely sociology. We also conducted a broad search in the JSTOR database to identify scholarly sociology articles published between 1998 and 2017 in English with a focus on defining or explaining qualitative research. We specifically zoom in on this time frame because we would have expect that this more mature period would have produced clear discussions on the meaning of qualitative research. To find these articles we combined a number of keywords to search the content and/or the title: qualitative (which was always included), definition, empirical, research, methodology, studies, fieldwork, interview and observation .

As a second phase of our research we searched within nine major sociological journals ( American Journal of Sociology , Sociological Theory , American Sociological Review , Contemporary Sociology , Sociological Forum , Sociological Theory , Qualitative Research , Qualitative Sociology and Qualitative Sociology Review ) for articles also published during the past 19 years (1998–2017) that had the term “qualitative” in the title and attempted to define qualitative research.

Lastly we picked two additional journals, Qualitative Research and Qualitative Sociology , in which we could expect to find texts addressing the notion of “qualitative.” From Qualitative Research we chose Volume 14, Issue 6, December 2014, and from Qualitative Sociology we chose Volume 36, Issue 2, June 2017. Within each of these we selected the first article; then we picked the second article of three prior issues. Again we went back another three issues and investigated article number three. Finally we went back another three issues and perused article number four. This selection criteria was used to get a manageable sample for the analysis.

The coding process of the 89 references we gathered in our selected review began soon after the first round of material was gathered, and we reduced the complexity created by our maximum variation sampling (Snow and Anderson 1993 :22) to four different categories within which questions on the nature and properties of qualitative research were discussed. We call them: Qualitative and Quantitative Research, Qualitative Research, Fieldwork, and Grounded Theory. This – which may appear as an illogical grouping – merely reflects the “context” in which the matter of “qualitative” is discussed. If the selection process of the material – books and articles – was informed by pre-knowledge, we used an inductive strategy to code the material. When studying our material, we identified four central notions related to “qualitative” that appear in various combinations in the literature which indicate what is the core of qualitative research. We have labeled them: “distinctions”, “process,” “closeness,” and “improved understanding.” During the research process the categories and notions were improved, refined, changed, and reordered. The coding ended when a sense of saturation in the material arose. In the presentation below all quotations and references come from our empirical material of texts on qualitative research.

Analysis – What is Qualitative Research?

In this section we describe the four categories we identified in the coding, how they differently discuss qualitative research, as well as their overall content. Some salient quotations are selected to represent the type of text sorted under each of the four categories. What we present are examples from the literature.

Qualitative and Quantitative

This analytic category comprises quotations comparing qualitative and quantitative research, a distinction that is frequently used (Brown 2010 :231); in effect this is a conceptual pair that structures the discussion and that may be associated with opposing interests. While the general goal of quantitative and qualitative research is the same – to understand the world better – their methodologies and focus in certain respects differ substantially (Becker 1966 :55). Quantity refers to that property of something that can be determined by measurement. In a dictionary of Statistics and Methodology we find that “(a) When referring to *variables, ‘qualitative’ is another term for *categorical or *nominal. (b) When speaking of kinds of research, ‘qualitative’ refers to studies of subjects that are hard to quantify, such as art history. Qualitative research tends to be a residual category for almost any kind of non-quantitative research” (Stiles 1998:183). But it should be obvious that one could employ a quantitative approach when studying, for example, art history.

The same dictionary states that quantitative is “said of variables or research that can be handled numerically, usually (too sharply) contrasted with *qualitative variables and research” (Stiles 1998:184). From a qualitative perspective “quantitative research” is about numbers and counting, and from a quantitative perspective qualitative research is everything that is not about numbers. But this does not say much about what is “qualitative.” If we turn to encyclopedias we find that in the 1932 edition of the Encyclopedia of the Social Sciences there is no mention of “qualitative.” In the Encyclopedia from 1968 we can read:

Qualitative Analysis. For methods of obtaining, analyzing, and describing data, see [the various entries:] CONTENT ANALYSIS; COUNTED DATA; EVALUATION RESEARCH, FIELD WORK; GRAPHIC PRESENTATION; HISTORIOGRAPHY, especially the article on THE RHETORIC OF HISTORY; INTERVIEWING; OBSERVATION; PERSONALITY MEASUREMENT; PROJECTIVE METHODS; PSYCHOANALYSIS, article on EXPERIMENTAL METHODS; SURVEY ANALYSIS, TABULAR PRESENTATION; TYPOLOGIES. (Vol. 13:225)

Some, like Alford, divide researchers into methodologists or, in his words, “quantitative and qualitative specialists” (Alford 1998 :12). Qualitative research uses a variety of methods, such as intensive interviews or in-depth analysis of historical materials, and it is concerned with a comprehensive account of some event or unit (King et al. 1994 :4). Like quantitative research it can be utilized to study a variety of issues, but it tends to focus on meanings and motivations that underlie cultural symbols, personal experiences, phenomena and detailed understanding of processes in the social world. In short, qualitative research centers on understanding processes, experiences, and the meanings people assign to things (Kalof et al. 2008 :79).

Others simply say that qualitative methods are inherently unscientific (Jovanović 2011 :19). Hood, for instance, argues that words are intrinsically less precise than numbers, and that they are therefore more prone to subjective analysis, leading to biased results (Hood 2006 :219). Qualitative methodologies have raised concerns over the limitations of quantitative templates (Brady et al. 2004 :4). Scholars such as King et al. ( 1994 ), for instance, argue that non-statistical research can produce more reliable results if researchers pay attention to the rules of scientific inference commonly stated in quantitative research. Also, researchers such as Becker ( 1966 :59; 1970 :42–43) have asserted that, if conducted properly, qualitative research and in particular ethnographic field methods, can lead to more accurate results than quantitative studies, in particular, survey research and laboratory experiments.

Some researchers, such as Kalof, Dan, and Dietz ( 2008 :79) claim that the boundaries between the two approaches are becoming blurred, and Small ( 2009 ) argues that currently much qualitative research (especially in North America) tries unsuccessfully and unnecessarily to emulate quantitative standards. For others, qualitative research tends to be more humanistic and discursive (King et al. 1994 :4). Ragin ( 1994 ), and similarly also Becker, ( 1996 :53), Marchel and Owens ( 2007 :303) think that the main distinction between the two styles is overstated and does not rest on the simple dichotomy of “numbers versus words” (Ragin 1994 :xii). Some claim that quantitative data can be utilized to discover associations, but in order to unveil cause and effect a complex research design involving the use of qualitative approaches needs to be devised (Gilbert 2009 :35). Consequently, qualitative data are useful for understanding the nuances lying beyond those processes as they unfold (Gilbert 2009 :35). Others contend that qualitative research is particularly well suited both to identify causality and to uncover fine descriptive distinctions (Fine and Hallett 2014 ; Lichterman and Isaac Reed 2014 ; Katz 2015 ).

There are other ways to separate these two traditions, including normative statements about what qualitative research should be (that is, better or worse than quantitative approaches, concerned with scientific approaches to societal change or vice versa; Snow and Morrill 1995 ; Denzin and Lincoln 2005 ), or whether it should develop falsifiable statements; Best 2004 ).

We propose that quantitative research is largely concerned with pre-determined variables (Small 2008 ); the analysis concerns the relations between variables. These categories are primarily not questioned in the study, only their frequency or degree, or the correlations between them (cf. Franzosi 2016 ). If a researcher studies wage differences between women and men, he or she works with given categories: x number of men are compared with y number of women, with a certain wage attributed to each person. The idea is not to move beyond the given categories of wage, men and women; they are the starting point as well as the end point, and undergo no “qualitative change.” Qualitative research, in contrast, investigates relations between categories that are themselves subject to change in the research process. Returning to Becker’s study ( 1963 ), we see that he questioned pre-dispositional theories of deviant behavior working with pre-determined variables such as an individual’s combination of personal qualities or emotional problems. His take, in contrast, was to understand marijuana consumption by developing “variables” as part of the investigation. Thereby he presented new variables, or as we would say today, theoretical concepts, but which are grounded in the empirical material.

Qualitative Research

This category contains quotations that refer to descriptions of qualitative research without making comparisons with quantitative research. Researchers such as Denzin and Lincoln, who have written a series of influential handbooks on qualitative methods (1994; Denzin and Lincoln 2003 ; 2005 ), citing Nelson et al. (1992:4), argue that because qualitative research is “interdisciplinary, transdisciplinary, and sometimes counterdisciplinary” it is difficult to derive one single definition of it (Jovanović 2011 :3). According to them, in fact, “the field” is “many things at the same time,” involving contradictions, tensions over its focus, methods, and how to derive interpretations and findings ( 2003 : 11). Similarly, others, such as Flick ( 2007 :ix–x) contend that agreeing on an accepted definition has increasingly become problematic, and that qualitative research has possibly matured different identities. However, Best holds that “the proliferation of many sorts of activities under the label of qualitative sociology threatens to confuse our discussions” ( 2004 :54). Atkinson’s position is more definite: “the current state of qualitative research and research methods is confused” ( 2005 :3–4).

Qualitative research is about interpretation (Blumer 1969 ; Strauss and Corbin 1998 ; Denzin and Lincoln 2003 ), or Verstehen [understanding] (Frankfort-Nachmias and Nachmias 1996 ). It is “multi-method,” involving the collection and use of a variety of empirical materials (Denzin and Lincoln 1998; Silverman 2013 ) and approaches (Silverman 2005 ; Flick 2007 ). It focuses not only on the objective nature of behavior but also on its subjective meanings: individuals’ own accounts of their attitudes, motivations, behavior (McIntyre 2005 :127; Creswell 2009 ), events and situations (Bryman 1989) – what people say and do in specific places and institutions (Goodwin and Horowitz 2002 :35–36) in social and temporal contexts (Morrill and Fine 1997). For this reason, following Weber ([1921-22] 1978), it can be described as an interpretative science (McIntyre 2005 :127). But could quantitative research also be concerned with these questions? Also, as pointed out below, does all qualitative research focus on subjective meaning, as some scholars suggest?

Others also distinguish qualitative research by claiming that it collects data using a naturalistic approach (Denzin and Lincoln 2005 :2; Creswell 2009 ), focusing on the meaning actors ascribe to their actions. But again, does all qualitative research need to be collected in situ? And does qualitative research have to be inherently concerned with meaning? Flick ( 2007 ), referring to Denzin and Lincoln ( 2005 ), mentions conversation analysis as an example of qualitative research that is not concerned with the meanings people bring to a situation, but rather with the formal organization of talk. Still others, such as Ragin ( 1994 :85), note that qualitative research is often (especially early on in the project, we would add) less structured than other kinds of social research – a characteristic connected to its flexibility and that can lead both to potentially better, but also worse results. But is this not a feature of this type of research, rather than a defining description of its essence? Wouldn’t this comment also apply, albeit to varying degrees, to quantitative research?

In addition, Strauss ( 2003 ), along with others, such as Alvesson and Kärreman ( 2011 :10–76), argue that qualitative researchers struggle to capture and represent complex phenomena partially because they tend to collect a large amount of data. While his analysis is correct at some points – “It is necessary to do detailed, intensive, microscopic examination of the data in order to bring out the amazing complexity of what lies in, behind, and beyond those data” (Strauss 2003 :10) – much of his analysis concerns the supposed focus of qualitative research and its challenges, rather than exactly what it is about. But even in this instance we would make a weak case arguing that these are strictly the defining features of qualitative research. Some researchers seem to focus on the approach or the methods used, or even on the way material is analyzed. Several researchers stress the naturalistic assumption of investigating the world, suggesting that meaning and interpretation appear to be a core matter of qualitative research.

We can also see that in this category there is no consensus about specific qualitative methods nor about qualitative data. Many emphasize interpretation, but quantitative research, too, involves interpretation; the results of a regression analysis, for example, certainly have to be interpreted, and the form of meta-analysis that factor analysis provides indeed requires interpretation However, there is no interpretation of quantitative raw data, i.e., numbers in tables. One common thread is that qualitative researchers have to get to grips with their data in order to understand what is being studied in great detail, irrespective of the type of empirical material that is being analyzed. This observation is connected to the fact that qualitative researchers routinely make several adjustments of focus and research design as their studies progress, in many cases until the very end of the project (Kalof et al. 2008 ). If you, like Becker, do not start out with a detailed theory, adjustments such as the emergence and refinement of research questions will occur during the research process. We have thus found a number of useful reflections about qualitative research scattered across different sources, but none of them effectively describe the defining characteristics of this approach.

Although qualitative research does not appear to be defined in terms of a specific method, it is certainly common that fieldwork, i.e., research that entails that the researcher spends considerable time in the field that is studied and use the knowledge gained as data, is seen as emblematic of or even identical to qualitative research. But because we understand that fieldwork tends to focus primarily on the collection and analysis of qualitative data, we expected to find within it discussions on the meaning of “qualitative.” But, again, this was not the case.

Instead, we found material on the history of this approach (for example, Frankfort-Nachmias and Nachmias 1996 ; Atkinson et al. 2001), including how it has changed; for example, by adopting a more self-reflexive practice (Heyl 2001), as well as the different nomenclature that has been adopted, such as fieldwork, ethnography, qualitative research, naturalistic research, participant observation and so on (for example, Lofland et al. 2006 ; Gans 1999 ).

We retrieved definitions of ethnography, such as “the study of people acting in the natural courses of their daily lives,” involving a “resocialization of the researcher” (Emerson 1988 :1) through intense immersion in others’ social worlds (see also examples in Hammersley 2018 ). This may be accomplished by direct observation and also participation (Neuman 2007 :276), although others, such as Denzin ( 1970 :185), have long recognized other types of observation, including non-participant (“fly on the wall”). In this category we have also isolated claims and opposing views, arguing that this type of research is distinguished primarily by where it is conducted (natural settings) (Hughes 1971:496), and how it is carried out (a variety of methods are applied) or, for some most importantly, by involving an active, empathetic immersion in those being studied (Emerson 1988 :2). We also retrieved descriptions of the goals it attends in relation to how it is taught (understanding subjective meanings of the people studied, primarily develop theory, or contribute to social change) (see for example, Corte and Irwin 2017 ; Frankfort-Nachmias and Nachmias 1996 :281; Trier-Bieniek 2012 :639) by collecting the richest possible data (Lofland et al. 2006 ) to derive “thick descriptions” (Geertz 1973 ), and/or to aim at theoretical statements of general scope and applicability (for example, Emerson 1988 ; Fine 2003 ). We have identified guidelines on how to evaluate it (for example Becker 1996 ; Lamont 2004 ) and have retrieved instructions on how it should be conducted (for example, Lofland et al. 2006 ). For instance, analysis should take place while the data gathering unfolds (Emerson 1988 ; Hammersley and Atkinson 2007 ; Lofland et al. 2006 ), observations should be of long duration (Becker 1970 :54; Goffman 1989 ), and data should be of high quantity (Becker 1970 :52–53), as well as other questionable distinctions between fieldwork and other methods:

Field studies differ from other methods of research in that the researcher performs the task of selecting topics, decides what questions to ask, and forges interest in the course of the research itself . This is in sharp contrast to many ‘theory-driven’ and ‘hypothesis-testing’ methods. (Lofland and Lofland 1995 :5)

But could not, for example, a strictly interview-based study be carried out with the same amount of flexibility, such as sequential interviewing (for example, Small 2009 )? Once again, are quantitative approaches really as inflexible as some qualitative researchers think? Moreover, this category stresses the role of the actors’ meaning, which requires knowledge and close interaction with people, their practices and their lifeworld.

It is clear that field studies – which are seen by some as the “gold standard” of qualitative research – are nonetheless only one way of doing qualitative research. There are other methods, but it is not clear why some are more qualitative than others, or why they are better or worse. Fieldwork is characterized by interaction with the field (the material) and understanding of the phenomenon that is being studied. In Becker’s case, he had general experience from fields in which marihuana was used, based on which he did interviews with actual users in several fields.

Grounded Theory

Another major category we identified in our sample is Grounded Theory. We found descriptions of it most clearly in Glaser and Strauss’ ([1967] 2010 ) original articulation, Strauss and Corbin ( 1998 ) and Charmaz ( 2006 ), as well as many other accounts of what it is for: generating and testing theory (Strauss 2003 :xi). We identified explanations of how this task can be accomplished – such as through two main procedures: constant comparison and theoretical sampling (Emerson 1998:96), and how using it has helped researchers to “think differently” (for example, Strauss and Corbin 1998 :1). We also read descriptions of its main traits, what it entails and fosters – for instance, an exceptional flexibility, an inductive approach (Strauss and Corbin 1998 :31–33; 1990; Esterberg 2002 :7), an ability to step back and critically analyze situations, recognize tendencies towards bias, think abstractly and be open to criticism, enhance sensitivity towards the words and actions of respondents, and develop a sense of absorption and devotion to the research process (Strauss and Corbin 1998 :5–6). Accordingly, we identified discussions of the value of triangulating different methods (both using and not using grounded theory), including quantitative ones, and theories to achieve theoretical development (most comprehensively in Denzin 1970 ; Strauss and Corbin 1998 ; Timmermans and Tavory 2012 ). We have also located arguments about how its practice helps to systematize data collection, analysis and presentation of results (Glaser and Strauss [1967] 2010 :16).

Grounded theory offers a systematic approach which requires researchers to get close to the field; closeness is a requirement of identifying questions and developing new concepts or making further distinctions with regard to old concepts. In contrast to other qualitative approaches, grounded theory emphasizes the detailed coding process, and the numerous fine-tuned distinctions that the researcher makes during the process. Within this category, too, we could not find a satisfying discussion of the meaning of qualitative research.

Defining Qualitative Research

In sum, our analysis shows that some notions reappear in the discussion of qualitative research, such as understanding, interpretation, “getting close” and making distinctions. These notions capture aspects of what we think is “qualitative.” However, a comprehensive definition that is useful and that can further develop the field is lacking, and not even a clear picture of its essential elements appears. In other words no definition emerges from our data, and in our research process we have moved back and forth between our empirical data and the attempt to present a definition. Our concrete strategy, as stated above, is to relate qualitative and quantitative research, or more specifically, qualitative and quantitative work. We use an ideal-typical notion of quantitative research which relies on taken for granted and numbered variables. This means that the data consists of variables on different scales, such as ordinal, but frequently ratio and absolute scales, and the representation of the numbers to the variables, i.e. the justification of the assignment of numbers to object or phenomenon, are not questioned, though the validity may be questioned. In this section we return to the notion of quality and try to clarify it while presenting our contribution.

Broadly, research refers to the activity performed by people trained to obtain knowledge through systematic procedures. Notions such as “objectivity” and “reflexivity,” “systematic,” “theory,” “evidence” and “openness” are here taken for granted in any type of research. Next, building on our empirical analysis we explain the four notions that we have identified as central to qualitative work: distinctions, process, closeness, and improved understanding. In discussing them, ultimately in relation to one another, we make their meaning even more precise. Our idea, in short, is that only when these ideas that we present separately for analytic purposes are brought together can we speak of qualitative research.

Distinctions

We believe that the possibility of making new distinctions is one the defining characteristics of qualitative research. It clearly sets it apart from quantitative analysis which works with taken-for-granted variables, albeit as mentioned, meta-analyses, for example, factor analysis may result in new variables. “Quality” refers essentially to distinctions, as already pointed out by Aristotle. He discusses the term “qualitative” commenting: “By a quality I mean that in virtue of which things are said to be qualified somehow” (Aristotle 1984:14). Quality is about what something is or has, which means that the distinction from its environment is crucial. We see qualitative research as a process in which significant new distinctions are made to the scholarly community; to make distinctions is a key aspect of obtaining new knowledge; a point, as we will see, that also has implications for “quantitative research.” The notion of being “significant” is paramount. New distinctions by themselves are not enough; just adding concepts only increases complexity without furthering our knowledge. The significance of new distinctions is judged against the communal knowledge of the research community. To enable this discussion and judgements central elements of rational discussion are required (cf. Habermas [1981] 1987 ; Davidsson [ 1988 ] 2001) to identify what is new and relevant scientific knowledge. Relatedly, Ragin alludes to the idea of new and useful knowledge at a more concrete level: “Qualitative methods are appropriate for in-depth examination of cases because they aid the identification of key features of cases. Most qualitative methods enhance data” (1994:79). When Becker ( 1963 ) studied deviant behavior and investigated how people became marihuana smokers, he made distinctions between the ways in which people learned how to smoke. This is a classic example of how the strategy of “getting close” to the material, for example the text, people or pictures that are subject to analysis, may enable researchers to obtain deeper insight and new knowledge by making distinctions – in this instance on the initial notion of learning how to smoke. Others have stressed the making of distinctions in relation to coding or theorizing. Emerson et al. ( 1995 ), for example, hold that “qualitative coding is a way of opening up avenues of inquiry,” meaning that the researcher identifies and develops concepts and analytic insights through close examination of and reflection on data (Emerson et al. 1995 :151). Goodwin and Horowitz highlight making distinctions in relation to theory-building writing: “Close engagement with their cases typically requires qualitative researchers to adapt existing theories or to make new conceptual distinctions or theoretical arguments to accommodate new data” ( 2002 : 37). In the ideal-typical quantitative research only existing and so to speak, given, variables would be used. If this is the case no new distinction are made. But, would not also many “quantitative” researchers make new distinctions?

Process does not merely suggest that research takes time. It mainly implies that qualitative new knowledge results from a process that involves several phases, and above all iteration. Qualitative research is about oscillation between theory and evidence, analysis and generating material, between first- and second -order constructs (Schütz 1962 :59), between getting in contact with something, finding sources, becoming deeply familiar with a topic, and then distilling and communicating some of its essential features. The main point is that the categories that the researcher uses, and perhaps takes for granted at the beginning of the research process, usually undergo qualitative changes resulting from what is found. Becker describes how he tested hypotheses and let the jargon of the users develop into theoretical concepts. This happens over time while the study is being conducted, exemplifying what we mean by process.

In the research process, a pilot-study may be used to get a first glance of, for example, the field, how to approach it, and what methods can be used, after which the method and theory are chosen or refined before the main study begins. Thus, the empirical material is often central from the start of the project and frequently leads to adjustments by the researcher. Likewise, during the main study categories are not fixed; the empirical material is seen in light of the theory used, but it is also given the opportunity to kick back, thereby resisting attempts to apply theoretical straightjackets (Becker 1970 :43). In this process, coding and analysis are interwoven, and thus are often important steps for getting closer to the phenomenon and deciding what to focus on next. Becker began his research by interviewing musicians close to him, then asking them to refer him to other musicians, and later on doubling his original sample of about 25 to include individuals in other professions (Becker 1973:46). Additionally, he made use of some participant observation, documents, and interviews with opiate users made available to him by colleagues. As his inductive theory of deviance evolved, Becker expanded his sample in order to fine tune it, and test the accuracy and generality of his hypotheses. In addition, he introduced a negative case and discussed the null hypothesis ( 1963 :44). His phasic career model is thus based on a research design that embraces processual work. Typically, process means to move between “theory” and “material” but also to deal with negative cases, and Becker ( 1998 ) describes how discovering these negative cases impacted his research design and ultimately its findings.

Obviously, all research is process-oriented to some degree. The point is that the ideal-typical quantitative process does not imply change of the data, and iteration between data, evidence, hypotheses, empirical work, and theory. The data, quantified variables, are, in most cases fixed. Merging of data, which of course can be done in a quantitative research process, does not mean new data. New hypotheses are frequently tested, but the “raw data is often the “the same.” Obviously, over time new datasets are made available and put into use.

Another characteristic that is emphasized in our sample is that qualitative researchers – and in particular ethnographers – can, or as Goffman put it, ought to ( 1989 ), get closer to the phenomenon being studied and their data than quantitative researchers (for example, Silverman 2009 :85). Put differently, essentially because of their methods qualitative researchers get into direct close contact with those being investigated and/or the material, such as texts, being analyzed. Becker started out his interview study, as we noted, by talking to those he knew in the field of music to get closer to the phenomenon he was studying. By conducting interviews he got even closer. Had he done more observations, he would undoubtedly have got even closer to the field.

Additionally, ethnographers’ design enables researchers to follow the field over time, and the research they do is almost by definition longitudinal, though the time in the field is studied obviously differs between studies. The general characteristic of closeness over time maximizes the chances of unexpected events, new data (related, for example, to archival research as additional sources, and for ethnography for situations not necessarily previously thought of as instrumental – what Mannay and Morgan ( 2015 ) term the “waiting field”), serendipity (Merton and Barber 2004 ; Åkerström 2013 ), and possibly reactivity, as well as the opportunity to observe disrupted patterns that translate into exemplars of negative cases. Two classic examples of this are Becker’s finding of what medical students call “crocks” (Becker et al. 1961 :317), and Geertz’s ( 1973 ) study of “deep play” in Balinese society.

By getting and staying so close to their data – be it pictures, text or humans interacting (Becker was himself a musician) – for a long time, as the research progressively focuses, qualitative researchers are prompted to continually test their hunches, presuppositions and hypotheses. They test them against a reality that often (but certainly not always), and practically, as well as metaphorically, talks back, whether by validating them, or disqualifying their premises – correctly, as well as incorrectly (Fine 2003 ; Becker 1970 ). This testing nonetheless often leads to new directions for the research. Becker, for example, says that he was initially reading psychological theories, but when facing the data he develops a theory that looks at, you may say, everything but psychological dispositions to explain the use of marihuana. Especially researchers involved with ethnographic methods have a fairly unique opportunity to dig up and then test (in a circular, continuous and temporal way) new research questions and findings as the research progresses, and thereby to derive previously unimagined and uncharted distinctions by getting closer to the phenomenon under study.

Let us stress that getting close is by no means restricted to ethnography. The notion of hermeneutic circle and hermeneutics as a general way of understanding implies that we must get close to the details in order to get the big picture. This also means that qualitative researchers can literally also make use of details of pictures as evidence (cf. Harper 2002). Thus, researchers may get closer both when generating the material or when analyzing it.

Quantitative research, we maintain, in the ideal-typical representation cannot get closer to the data. The data is essentially numbers in tables making up the variables (Franzosi 2016 :138). The data may originally have been “qualitative,” but once reduced to numbers there can only be a type of “hermeneutics” about what the number may stand for. The numbers themselves, however, are non-ambiguous. Thus, in quantitative research, interpretation, if done, is not about the data itself—the numbers—but what the numbers stand for. It follows that the interpretation is essentially done in a more “speculative” mode without direct empirical evidence (cf. Becker 2017 ).

Improved Understanding

While distinction, process and getting closer refer to the qualitative work of the researcher, improved understanding refers to its conditions and outcome of this work. Understanding cuts deeper than explanation, which to some may mean a causally verified correlation between variables. The notion of explanation presupposes the notion of understanding since explanation does not include an idea of how knowledge is gained (Manicas 2006 : 15). Understanding, we argue, is the core concept of what we call the outcome of the process when research has made use of all the other elements that were integrated in the research. Understanding, then, has a special status in qualitative research since it refers both to the conditions of knowledge and the outcome of the process. Understanding can to some extent be seen as the condition of explanation and occurs in a process of interpretation, which naturally refers to meaning (Gadamer 1990 ). It is fundamentally connected to knowing, and to the knowing of how to do things (Heidegger [1927] 2001 ). Conceptually the term hermeneutics is used to account for this process. Heidegger ties hermeneutics to human being and not possible to separate from the understanding of being ( 1988 ). Here we use it in a broader sense, and more connected to method in general (cf. Seiffert 1992 ). The abovementioned aspects – for example, “objectivity” and “reflexivity” – of the approach are conditions of scientific understanding. Understanding is the result of a circular process and means that the parts are understood in light of the whole, and vice versa. Understanding presupposes pre-understanding, or in other words, some knowledge of the phenomenon studied. The pre-understanding, even in the form of prejudices, are in qualitative research process, which we see as iterative, questioned, which gradually or suddenly change due to the iteration of data, evidence and concepts. However, qualitative research generates understanding in the iterative process when the researcher gets closer to the data, e.g., by going back and forth between field and analysis in a process that generates new data that changes the evidence, and, ultimately, the findings. Questioning, to ask questions, and put what one assumes—prejudices and presumption—in question, is central to understand something (Heidegger [1927] 2001 ; Gadamer 1990 :368–384). We propose that this iterative process in which the process of understanding occurs is characteristic of qualitative research.

Improved understanding means that we obtain scientific knowledge of something that we as a scholarly community did not know before, or that we get to know something better. It means that we understand more about how parts are related to one another, and to other things we already understand (see also Fine and Hallett 2014 ). Understanding is an important condition for qualitative research. It is not enough to identify correlations, make distinctions, and work in a process in which one gets close to the field or phenomena. Understanding is accomplished when the elements are integrated in an iterative process.

It is, moreover, possible to understand many things, and researchers, just like children, may come to understand new things every day as they engage with the world. This subjective condition of understanding – namely, that a person gains a better understanding of something –is easily met. To be qualified as “scientific,” the understanding must be general and useful to many; it must be public. But even this generally accessible understanding is not enough in order to speak of “scientific understanding.” Though we as a collective can increase understanding of everything in virtually all potential directions as a result also of qualitative work, we refrain from this “objective” way of understanding, which has no means of discriminating between what we gain in understanding. Scientific understanding means that it is deemed relevant from the scientific horizon (compare Schütz 1962 : 35–38, 46, 63), and that it rests on the pre-understanding that the scientists have and must have in order to understand. In other words, the understanding gained must be deemed useful by other researchers, so that they can build on it. We thus see understanding from a pragmatic, rather than a subjective or objective perspective. Improved understanding is related to the question(s) at hand. Understanding, in order to represent an improvement, must be an improvement in relation to the existing body of knowledge of the scientific community (James [ 1907 ] 1955). Scientific understanding is, by definition, collective, as expressed in Weber’s famous note on objectivity, namely that scientific work aims at truths “which … can claim, even for a Chinese, the validity appropriate to an empirical analysis” ([1904] 1949 :59). By qualifying “improved understanding” we argue that it is a general defining characteristic of qualitative research. Becker‘s ( 1966 ) study and other research of deviant behavior increased our understanding of the social learning processes of how individuals start a behavior. And it also added new knowledge about the labeling of deviant behavior as a social process. Few studies, of course, make the same large contribution as Becker’s, but are nonetheless qualitative research.

Understanding in the phenomenological sense, which is a hallmark of qualitative research, we argue, requires meaning and this meaning is derived from the context, and above all the data being analyzed. The ideal-typical quantitative research operates with given variables with different numbers. This type of material is not enough to establish meaning at the level that truly justifies understanding. In other words, many social science explanations offer ideas about correlations or even causal relations, but this does not mean that the meaning at the level of the data analyzed, is understood. This leads us to say that there are indeed many explanations that meet the criteria of understanding, for example the explanation of how one becomes a marihuana smoker presented by Becker. However, we may also understand a phenomenon without explaining it, and we may have potential explanations, or better correlations, that are not really understood.

We may speak more generally of quantitative research and its data to clarify what we see as an important distinction. The “raw data” that quantitative research—as an idealtypical activity, refers to is not available for further analysis; the numbers, once created, are not to be questioned (Franzosi 2016 : 138). If the researcher is to do “more” or “change” something, this will be done by conjectures based on theoretical knowledge or based on the researcher’s lifeworld. Both qualitative and quantitative research is based on the lifeworld, and all researchers use prejudices and pre-understanding in the research process. This idea is present in the works of Heidegger ( 2001 ) and Heisenberg (cited in Franzosi 2010 :619). Qualitative research, as we argued, involves the interaction and questioning of concepts (theory), data, and evidence.

Ragin ( 2004 :22) points out that “a good definition of qualitative research should be inclusive and should emphasize its key strengths and features, not what it lacks (for example, the use of sophisticated quantitative techniques).” We define qualitative research as an iterative process in which improved understanding to the scientific community is achieved by making new significant distinctions resulting from getting closer to the phenomenon studied. Qualitative research, as defined here, is consequently a combination of two criteria: (i) how to do things –namely, generating and analyzing empirical material, in an iterative process in which one gets closer by making distinctions, and (ii) the outcome –improved understanding novel to the scholarly community. Is our definition applicable to our own study? In this study we have closely read the empirical material that we generated, and the novel distinction of the notion “qualitative research” is the outcome of an iterative process in which both deduction and induction were involved, in which we identified the categories that we analyzed. We thus claim to meet the first criteria, “how to do things.” The second criteria cannot be judged but in a partial way by us, namely that the “outcome” —in concrete form the definition-improves our understanding to others in the scientific community.

We have defined qualitative research, or qualitative scientific work, in relation to quantitative scientific work. Given this definition, qualitative research is about questioning the pre-given (taken for granted) variables, but it is thus also about making new distinctions of any type of phenomenon, for example, by coining new concepts, including the identification of new variables. This process, as we have discussed, is carried out in relation to empirical material, previous research, and thus in relation to theory. Theory and previous research cannot be escaped or bracketed. According to hermeneutic principles all scientific work is grounded in the lifeworld, and as social scientists we can thus never fully bracket our pre-understanding.

We have proposed that quantitative research, as an idealtype, is concerned with pre-determined variables (Small 2008 ). Variables are epistemically fixed, but can vary in terms of dimensions, such as frequency or number. Age is an example; as a variable it can take on different numbers. In relation to quantitative research, qualitative research does not reduce its material to number and variables. If this is done the process of comes to a halt, the researcher gets more distanced from her data, and it makes it no longer possible to make new distinctions that increase our understanding. We have above discussed the components of our definition in relation to quantitative research. Our conclusion is that in the research that is called quantitative there are frequent and necessary qualitative elements.

Further, comparative empirical research on researchers primarily working with ”quantitative” approaches and those working with ”qualitative” approaches, we propose, would perhaps show that there are many similarities in practices of these two approaches. This is not to deny dissimilarities, or the different epistemic and ontic presuppositions that may be more or less strongly associated with the two different strands (see Goertz and Mahoney 2012 ). Our point is nonetheless that prejudices and preconceptions about researchers are unproductive, and that as other researchers have argued, differences may be exaggerated (e.g., Becker 1996 : 53, 2017 ; Marchel and Owens 2007 :303; Ragin 1994 ), and that a qualitative dimension is present in both kinds of work.

Several things follow from our findings. The most important result is the relation to quantitative research. In our analysis we have separated qualitative research from quantitative research. The point is not to label individual researchers, methods, projects, or works as either “quantitative” or “qualitative.” By analyzing, i.e., taking apart, the notions of quantitative and qualitative, we hope to have shown the elements of qualitative research. Our definition captures the elements, and how they, when combined in practice, generate understanding. As many of the quotations we have used suggest, one conclusion of our study holds that qualitative approaches are not inherently connected with a specific method. Put differently, none of the methods that are frequently labelled “qualitative,” such as interviews or participant observation, are inherently “qualitative.” What matters, given our definition, is whether one works qualitatively or quantitatively in the research process, until the results are produced. Consequently, our analysis also suggests that those researchers working with what in the literature and in jargon is often called “quantitative research” are almost bound to make use of what we have identified as qualitative elements in any research project. Our findings also suggest that many” quantitative” researchers, at least to some extent, are engaged with qualitative work, such as when research questions are developed, variables are constructed and combined, and hypotheses are formulated. Furthermore, a research project may hover between “qualitative” and “quantitative” or start out as “qualitative” and later move into a “quantitative” (a distinct strategy that is not similar to “mixed methods” or just simply combining induction and deduction). More generally speaking, the categories of “qualitative” and “quantitative,” unfortunately, often cover up practices, and it may lead to “camps” of researchers opposing one another. For example, regardless of the researcher is primarily oriented to “quantitative” or “qualitative” research, the role of theory is neglected (cf. Swedberg 2017 ). Our results open up for an interaction not characterized by differences, but by different emphasis, and similarities.

Let us take two examples to briefly indicate how qualitative elements can fruitfully be combined with quantitative. Franzosi ( 2010 ) has discussed the relations between quantitative and qualitative approaches, and more specifically the relation between words and numbers. He analyzes texts and argues that scientific meaning cannot be reduced to numbers. Put differently, the meaning of the numbers is to be understood by what is taken for granted, and what is part of the lifeworld (Schütz 1962 ). Franzosi shows how one can go about using qualitative and quantitative methods and data to address scientific questions analyzing violence in Italy at the time when fascism was rising (1919–1922). Aspers ( 2006 ) studied the meaning of fashion photographers. He uses an empirical phenomenological approach, and establishes meaning at the level of actors. In a second step this meaning, and the different ideal-typical photographers constructed as a result of participant observation and interviews, are tested using quantitative data from a database; in the first phase to verify the different ideal-types, in the second phase to use these types to establish new knowledge about the types. In both of these cases—and more examples can be found—authors move from qualitative data and try to keep the meaning established when using the quantitative data.

A second main result of our study is that a definition, and we provided one, offers a way for research to clarify, and even evaluate, what is done. Hence, our definition can guide researchers and students, informing them on how to think about concrete research problems they face, and to show what it means to get closer in a process in which new distinctions are made. The definition can also be used to evaluate the results, given that it is a standard of evaluation (cf. Hammersley 2007 ), to see whether new distinctions are made and whether this improves our understanding of what is researched, in addition to the evaluation of how the research was conducted. By making what is qualitative research explicit it becomes easier to communicate findings, and it is thereby much harder to fly under the radar with substandard research since there are standards of evaluation which make it easier to separate “good” from “not so good” qualitative research.

To conclude, our analysis, which ends with a definition of qualitative research can thus both address the “internal” issues of what is qualitative research, and the “external” critiques that make it harder to do qualitative research, to which both pressure from quantitative methods and general changes in society contribute.

Åkerström, Malin. 2013. Curiosity and serendipity in qualitative research. Qualitative Sociology Review 9 (2): 10–18.

Google Scholar  

Alford, Robert R. 1998. The craft of inquiry. Theories, methods, evidence . Oxford: Oxford University Press.

Alvesson, Mats, and Dan Kärreman. 2011. Qualitative research and theory development. Mystery as method . London: SAGE Publications.

Book   Google Scholar  

Aspers, Patrik. 2006. Markets in Fashion, A Phenomenological Approach. London Routledge.

Atkinson, Paul. 2005. Qualitative research. Unity and diversity. Forum: Qualitative Social Research 6 (3): 1–15.

Becker, Howard S. 1963. Outsiders. Studies in the sociology of deviance . New York: The Free Press.

Becker, Howard S. 1966. Whose side are we on? Social Problems 14 (3): 239–247.

Article   Google Scholar  

Becker, Howard S. 1970. Sociological work. Method and substance . New Brunswick: Transaction Books.

Becker, Howard S. 1996. The epistemology of qualitative research. In Ethnography and human development. Context and meaning in social inquiry , ed. Jessor Richard, Colby Anne, and Richard A. Shweder, 53–71. Chicago: University of Chicago Press.

Becker, Howard S. 1998. Tricks of the trade. How to think about your research while you're doing it . Chicago: University of Chicago Press.

Becker, Howard S. 2017. Evidence . Chigaco: University of Chicago Press.

Becker, Howard, Blanche Geer, Everett Hughes, and Anselm Strauss. 1961. Boys in White, student culture in medical school . New Brunswick: Transaction Publishers.

Berezin, Mabel. 2014. How do we know what we mean? Epistemological dilemmas in cultural sociology. Qualitative Sociology 37 (2): 141–151.

Best, Joel. 2004. Defining qualitative research. In Workshop on Scientific Foundations of Qualitative Research , eds . Charles, Ragin, Joanne, Nagel, and Patricia White, 53-54. http://www.nsf.gov/pubs/2004/nsf04219/nsf04219.pdf .

Biernacki, Richard. 2014. Humanist interpretation versus coding text samples. Qualitative Sociology 37 (2): 173–188.

Blumer, Herbert. 1969. Symbolic interactionism: Perspective and method . Berkeley: University of California Press.

Brady, Henry, David Collier, and Jason Seawright. 2004. Refocusing the discussion of methodology. In Rethinking social inquiry. Diverse tools, shared standards , ed. Brady Henry and Collier David, 3–22. Lanham: Rowman and Littlefield.

Brown, Allison P. 2010. Qualitative method and compromise in applied social research. Qualitative Research 10 (2): 229–248.

Charmaz, Kathy. 2006. Constructing grounded theory . London: Sage.

Corte, Ugo, and Katherine Irwin. 2017. “The Form and Flow of Teaching Ethnographic Knowledge: Hands-on Approaches for Learning Epistemology” Teaching Sociology 45(3): 209-219.

Creswell, John W. 2009. Research design. Qualitative, quantitative, and mixed method approaches . 3rd ed. Thousand Oaks: SAGE Publications.

Davidsson, David. 1988. 2001. The myth of the subjective. In Subjective, intersubjective, objective , ed. David Davidsson, 39–52. Oxford: Oxford University Press.

Denzin, Norman K. 1970. The research act: A theoretical introduction to Ssociological methods . Chicago: Aldine Publishing Company Publishers.

Denzin, Norman K., and Yvonna S. Lincoln. 2003. Introduction. The discipline and practice of qualitative research. In Collecting and interpreting qualitative materials , ed. Norman K. Denzin and Yvonna S. Lincoln, 1–45. Thousand Oaks: SAGE Publications.

Denzin, Norman K., and Yvonna S. Lincoln. 2005. Introduction. The discipline and practice of qualitative research. In The Sage handbook of qualitative research , ed. Norman K. Denzin and Yvonna S. Lincoln, 1–32. Thousand Oaks: SAGE Publications.

Emerson, Robert M., ed. 1988. Contemporary field research. A collection of readings . Prospect Heights: Waveland Press.

Emerson, Robert M., Rachel I. Fretz, and Linda L. Shaw. 1995. Writing ethnographic fieldnotes . Chicago: University of Chicago Press.

Esterberg, Kristin G. 2002. Qualitative methods in social research . Boston: McGraw-Hill.

Fine, Gary Alan. 1995. Review of “handbook of qualitative research.” Contemporary Sociology 24 (3): 416–418.

Fine, Gary Alan. 2003. “ Toward a Peopled Ethnography: Developing Theory from Group Life.” Ethnography . 4(1):41-60.

Fine, Gary Alan, and Black Hawk Hancock. 2017. The new ethnographer at work. Qualitative Research 17 (2): 260–268.

Fine, Gary Alan, and Timothy Hallett. 2014. Stranger and stranger: Creating theory through ethnographic distance and authority. Journal of Organizational Ethnography 3 (2): 188–203.

Flick, Uwe. 2002. Qualitative research. State of the art. Social Science Information 41 (1): 5–24.

Flick, Uwe. 2007. Designing qualitative research . London: SAGE Publications.

Frankfort-Nachmias, Chava, and David Nachmias. 1996. Research methods in the social sciences . 5th ed. London: Edward Arnold.

Franzosi, Roberto. 2010. Sociology, narrative, and the quality versus quantity debate (Goethe versus Newton): Can computer-assisted story grammars help us understand the rise of Italian fascism (1919- 1922)? Theory and Society 39 (6): 593–629.

Franzosi, Roberto. 2016. From method and measurement to narrative and number. International journal of social research methodology 19 (1): 137–141.

Gadamer, Hans-Georg. 1990. Wahrheit und Methode, Grundzüge einer philosophischen Hermeneutik . Band 1, Hermeneutik. Tübingen: J.C.B. Mohr.

Gans, Herbert. 1999. Participant Observation in an Age of “Ethnography”. Journal of Contemporary Ethnography 28 (5): 540–548.

Geertz, Clifford. 1973. The interpretation of cultures . New York: Basic Books.

Gilbert, Nigel. 2009. Researching social life . 3rd ed. London: SAGE Publications.

Glaeser, Andreas. 2014. Hermeneutic institutionalism: Towards a new synthesis. Qualitative Sociology 37: 207–241.

Glaser, Barney G., and Anselm L. Strauss. [1967] 2010. The discovery of grounded theory. Strategies for qualitative research. Hawthorne: Aldine.

Goertz, Gary, and James Mahoney. 2012. A tale of two cultures: Qualitative and quantitative research in the social sciences . Princeton: Princeton University Press.

Goffman, Erving. 1989. On fieldwork. Journal of Contemporary Ethnography 18 (2): 123–132.

Goodwin, Jeff, and Ruth Horowitz. 2002. Introduction. The methodological strengths and dilemmas of qualitative sociology. Qualitative Sociology 25 (1): 33–47.

Habermas, Jürgen. [1981] 1987. The theory of communicative action . Oxford: Polity Press.

Hammersley, Martyn. 2007. The issue of quality in qualitative research. International Journal of Research & Method in Education 30 (3): 287–305.

Hammersley, Martyn. 2013. What is qualitative research? Bloomsbury Publishing.

Hammersley, Martyn. 2018. What is ethnography? Can it survive should it? Ethnography and Education 13 (1): 1–17.

Hammersley, Martyn, and Paul Atkinson. 2007. Ethnography. Principles in practice . London: Tavistock Publications.

Heidegger, Martin. [1927] 2001. Sein und Zeit . Tübingen: Max Niemeyer Verlag.

Heidegger, Martin. 1988. 1923. Ontologie. Hermeneutik der Faktizität, Gesamtausgabe II. Abteilung: Vorlesungen 1919-1944, Band 63, Frankfurt am Main: Vittorio Klostermann.

Hempel, Carl G. 1966. Philosophy of the natural sciences . Upper Saddle River: Prentice Hall.

Hood, Jane C. 2006. Teaching against the text. The case of qualitative methods. Teaching Sociology 34 (3): 207–223.

James, William. 1907. 1955. Pragmatism . New York: Meredian Books.

Jovanović, Gordana. 2011. Toward a social history of qualitative research. History of the Human Sciences 24 (2): 1–27.

Kalof, Linda, Amy Dan, and Thomas Dietz. 2008. Essentials of social research . London: Open University Press.

Katz, Jack. 2015. Situational evidence: Strategies for causal reasoning from observational field notes. Sociological Methods & Research 44 (1): 108–144.

King, Gary, Robert O. Keohane, S. Sidney, and S. Verba. 1994. Designing social inquiry. In Scientific inference in qualitative research . Princeton: Princeton University Press.

Chapter   Google Scholar  

Lamont, Michelle. 2004. Evaluating qualitative research: Some empirical findings and an agenda. In Report from workshop on interdisciplinary standards for systematic qualitative research , ed. M. Lamont and P. White, 91–95. Washington, DC: National Science Foundation.

Lamont, Michèle, and Ann Swidler. 2014. Methodological pluralism and the possibilities and limits of interviewing. Qualitative Sociology 37 (2): 153–171.

Lazarsfeld, Paul, and Alan Barton. 1982. Some functions of qualitative analysis in social research. In The varied sociology of Paul Lazarsfeld , ed. Patricia Kendall, 239–285. New York: Columbia University Press.

Lichterman, Paul, and Isaac Reed I (2014), Theory and Contrastive Explanation in Ethnography. Sociological methods and research. Prepublished 27 October 2014; https://doi.org/10.1177/0049124114554458 .

Lofland, John, and Lyn Lofland. 1995. Analyzing social settings. A guide to qualitative observation and analysis . 3rd ed. Belmont: Wadsworth.

Lofland, John, David A. Snow, Leon Anderson, and Lyn H. Lofland. 2006. Analyzing social settings. A guide to qualitative observation and analysis . 4th ed. Belmont: Wadsworth/Thomson Learning.

Long, Adrew F., and Mary Godfrey. 2004. An evaluation tool to assess the quality of qualitative research studies. International Journal of Social Research Methodology 7 (2): 181–196.

Lundberg, George. 1951. Social research: A study in methods of gathering data . New York: Longmans, Green and Co..

Malinowski, Bronislaw. 1922. Argonauts of the Western Pacific: An account of native Enterprise and adventure in the archipelagoes of Melanesian New Guinea . London: Routledge.

Manicas, Peter. 2006. A realist philosophy of science: Explanation and understanding . Cambridge: Cambridge University Press.

Marchel, Carol, and Stephanie Owens. 2007. Qualitative research in psychology. Could William James get a job? History of Psychology 10 (4): 301–324.

McIntyre, Lisa J. 2005. Need to know. Social science research methods . Boston: McGraw-Hill.

Merton, Robert K., and Elinor Barber. 2004. The travels and adventures of serendipity. A Study in Sociological Semantics and the Sociology of Science . Princeton: Princeton University Press.

Mannay, Dawn, and Melanie Morgan. 2015. Doing ethnography or applying a qualitative technique? Reflections from the ‘waiting field‘. Qualitative Research 15 (2): 166–182.

Neuman, Lawrence W. 2007. Basics of social research. Qualitative and quantitative approaches . 2nd ed. Boston: Pearson Education.

Ragin, Charles C. 1994. Constructing social research. The unity and diversity of method . Thousand Oaks: Pine Forge Press.

Ragin, Charles C. 2004. Introduction to session 1: Defining qualitative research. In Workshop on Scientific Foundations of Qualitative Research , 22, ed. Charles C. Ragin, Joane Nagel, Patricia White. http://www.nsf.gov/pubs/2004/nsf04219/nsf04219.pdf

Rawls, Anne. 2018. The Wartime narrative in US sociology, 1940–7: Stigmatizing qualitative sociology in the name of ‘science,’ European Journal of Social Theory (Online first).

Schütz, Alfred. 1962. Collected papers I: The problem of social reality . The Hague: Nijhoff.

Seiffert, Helmut. 1992. Einführung in die Hermeneutik . Tübingen: Franke.

Silverman, David. 2005. Doing qualitative research. A practical handbook . 2nd ed. London: SAGE Publications.

Silverman, David. 2009. A very short, fairly interesting and reasonably cheap book about qualitative research . London: SAGE Publications.

Silverman, David. 2013. What counts as qualitative research? Some cautionary comments. Qualitative Sociology Review 9 (2): 48–55.

Small, Mario L. 2009. “How many cases do I need?” on science and the logic of case selection in field-based research. Ethnography 10 (1): 5–38.

Small, Mario L 2008. Lost in translation: How not to make qualitative research more scientific. In Workshop on interdisciplinary standards for systematic qualitative research, ed in Michelle Lamont, and Patricia White, 165–171. Washington, DC: National Science Foundation.

Snow, David A., and Leon Anderson. 1993. Down on their luck: A study of homeless street people . Berkeley: University of California Press.

Snow, David A., and Calvin Morrill. 1995. New ethnographies: Review symposium: A revolutionary handbook or a handbook for revolution? Journal of Contemporary Ethnography 24 (3): 341–349.

Strauss, Anselm L. 2003. Qualitative analysis for social scientists . 14th ed. Chicago: Cambridge University Press.

Strauss, Anselm L., and Juliette M. Corbin. 1998. Basics of qualitative research. Techniques and procedures for developing grounded theory . 2nd ed. Thousand Oaks: Sage Publications.

Swedberg, Richard. 2017. Theorizing in sociological research: A new perspective, a new departure? Annual Review of Sociology 43: 189–206.

Swedberg, Richard. 1990. The new 'Battle of Methods'. Challenge January–February 3 (1): 33–38.

Timmermans, Stefan, and Iddo Tavory. 2012. Theory construction in qualitative research: From grounded theory to abductive analysis. Sociological Theory 30 (3): 167–186.

Trier-Bieniek, Adrienne. 2012. Framing the telephone interview as a participant-centred tool for qualitative research. A methodological discussion. Qualitative Research 12 (6): 630–644.

Valsiner, Jaan. 2000. Data as representations. Contextualizing qualitative and quantitative research strategies. Social Science Information 39 (1): 99–113.

Weber, Max. 1904. 1949. Objectivity’ in social Science and social policy. Ed. Edward A. Shils and Henry A. Finch, 49–112. New York: The Free Press.

Download references

Acknowledgements

Financial Support for this research is given by the European Research Council, CEV (263699). The authors are grateful to Susann Krieglsteiner for assistance in collecting the data. The paper has benefitted from the many useful comments by the three reviewers and the editor, comments by members of the Uppsala Laboratory of Economic Sociology, as well as Jukka Gronow, Sebastian Kohl, Marcin Serafin, Richard Swedberg, Anders Vassenden and Turid Rødne.

Author information

Authors and affiliations.

Department of Sociology, Uppsala University, Uppsala, Sweden

Patrik Aspers

Seminar for Sociology, Universität St. Gallen, St. Gallen, Switzerland

Department of Media and Social Sciences, University of Stavanger, Stavanger, Norway

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Patrik Aspers .

Additional information

Publisher’s note.

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

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Aspers, P., Corte, U. What is Qualitative in Qualitative Research. Qual Sociol 42 , 139–160 (2019). https://doi.org/10.1007/s11133-019-9413-7

Download citation

Published : 27 February 2019

Issue Date : 01 June 2019

DOI : https://doi.org/10.1007/s11133-019-9413-7

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Qualitative research
  • Epistemology
  • Philosophy of science
  • Phenomenology
  • Find a journal
  • Publish with us
  • Track your research

Perceptions of burnout among public sector physicians in Sierra Leone: A qualitative study

PLOS

  • This person is not on ResearchGate, or hasn't claimed this research yet.

Asad Naveed

Abstract and Figures

qualitative meaning of research

Discover the world's research

  • 25+ million members
  • 160+ million publication pages
  • 2.3+ billion citations

Sarah Adamson-Grimmer

  • Jason P Sullivan

Charles Czeisler

  • Int J Environ Res Publ Health

Reham Shalaby

  • Erica Mitchell Smith

Angela Keniston

  • Christine Cara Welles

Marisha Burden

  • Zione Banda
  • Mirriam Simbota

Chimwemwe Mula

  • Giuseppa Maresca

Francesco Corallo

  • Giulia Catanese

Viviana Lo Buono

  • Recruit researchers
  • Join for free
  • Login Email Tip: Most researchers use their institutional email address as their ResearchGate login Password Forgot password? Keep me logged in Log in or Continue with Google Welcome back! Please log in. Email · Hint Tip: Most researchers use their institutional email address as their ResearchGate login Password Forgot password? Keep me logged in Log in or Continue with Google No account? Sign up
  • Open access
  • Published: 16 September 2024

“ You close the door , wipe your sadness and put on a smiling face ”: a qualitative study of the emotional labour of healthcare professionals providing palliative care in nursing homes in France

  • Benoite Umubyeyi 1 ,
  • Danièle Leboul 1 &
  • Emmanuel Bagaragaza 1  

BMC Health Services Research volume  24 , Article number:  1070 ( 2024 ) Cite this article

Metrics details

Palliative care provided to frail and dying older persons in nursing homes results in intense emotions for residents and their relatives as well as for healthcare professionals. In France, scant attention has been given to how nursing home professionals manage their emotions when providing palliative care. This study analysed the emotional demands of providing palliative care in the nursing home context, the emotional strategies used by healthcare professionals to navigate such demands, and how these demands affect their emotional wellbeing.

This qualitative study used a multiple case study approach. We purposively selected nine nursing homes from three geographical provinces in France with diverse ownership statuses (public, private, associative). Individual interviews and focus group discussions were held with 93 healthcare professionals from various occupational groups employed in the participating nursing homes. Data was collected from April 2021 to September 2022 and was analysed using thematic content analysis.

Data revealed that providing palliative care to dying residents within the nursing home context results in intertwined rewarding and exhausting emotional experiences for healthcare professionals. Professionals have to utilize multifaceted emotional strategies to navigate these experiences, including suppressing and modifying emotions and distancing themselves emotionally from residents to protect themselves from emotional suffering. Participants noted a lack of formal space to express emotions. Unrecognized emotional labour undermines the wellbeing of healthcare professionals in nursing homes, whereas acknowledging emotions enhances satisfaction and gives enhanced meaning to their crucial role in resident care.

Acknowledging emotional labour as an inevitable component of providing palliative care in nursing homes is critical to supporting healthcare professional wellbeing, resilience, and retention, which may ultimately improve the quality of care for dying residents. Ensuring quality care and supporting the emotional wellbeing of nursing home professionals requires an organisational culture that considers emotional expression a collective strength-building resource rather than an individual responsibility, in hopes of shaping a new culture that fully acknowledges their humanity alongside their professional skills.

Trial registration

ClinicalTrials.gov ID: NCT04708002; National registration: ID-RCB number: 2020-A01832-37, Registration date: 2020-12-03.

Peer Review reports

The steady increase in the number of older persons affected by multiple and complex health needs has led to a growing number of nursing home (NH) residents worldwide spending their final moments of life and dying within these settings [ 1 ]. In turn, this trend has generated heightened attention for the necessity of integrating palliative care within the NH, an approach that has not traditionally been an area of focus [ 2 , 3 , 4 ].

The core philosophy and values of the NH are to provide a home-like environment for residents. When residents require palliative and end-of-life care, the focus shifts from supporting quality living to facilitating quality dying [ 5 ]. Such a shift is accompanied by intense emotions for residents, their relatives, and NH professionals who have had intimate interactions with residents and built strong ties and long-lasting relationships with them during the extended caregiving process. The process becomes even more emotionally laden for professionals, as they often see themselves as holding a professional caring role while also taking on the emotional work of a family member role [ 6 ]. When confronted with providing quality living while simultaneously supporting quality dying [ 5 ], NH professionals have to perform significant emotional labour to provide quality care and preserve the professional-resident therapeutic relationship, all the while maintaining their own health and emotional wellbeing [ 7 ].

In France, as in many other countries, the provision of palliative care in NHs relies heavily on a multidisciplinary staff mainly composed of nursing assistants, personal support workers, registered nurses, and other regulated professionals, under the supervision of a medical coordinator. A medical coordinator in the French nursing home context is a physician, generally with geriatric competences, who has an overall coordination and medical advisory role for nursing home and external provider team for enhanced quality care. Moreover, as elsewhere, NHs in France suffer significant staff shortages due to professional fatigue, burnout, and professional turnover [ 8 ]. The recent COVID-19 pandemic has worsened the situation in NHs, with increased COVID-19-related deaths, augmented workloads, expanded isolation, and added psychological burden among care workers [ 9 , 10 ]. There is a pressing need for NH organisations to acknowledge the emotional labour endured by healthcare professionals during the caregiving process, particularly when it involves providing palliative and end-of-life care to residents.

Current evidence has demonstrated a strong link between burnout, job satisfaction, performance, staff retention, and attrition and the emotional labour of caring [ 11 , 12 ]. When healthcare professionals have to suppress or modify their emotions, they experience dissonance between feelings and performance, which in the long term can result in emotional distress, burnout, and intention to leave the profession [ 12 , 13 ]. Other studies have noted that when emotions are freely expressed and supported, they may have a positive impact on professional-patient interpersonal relationships, staff member wellbeing, and the quality of patient care [ 14 ]. However, most studies that have explored the emotional labour involved in providing palliative care have focused on hospital, hospice, or palliative care unit settings. Rarely have these studies been conducted in NH contexts. In fact, the physical labor associated with caring in NHs and the economic aspects of the work, such as wages and scheduling, receive more attention than does emotional labour [ 15 ]. In addition, emotional intelligence is an expected competency of healthcare professionals, particularly an individual’s ability to manage their own emotions to the point that failing to do so is viewed as an individual weakness and professional failure [ 16 ]. Yet, it has been documented that the effectiveness of emotional labour depends on various factors, including the nature of the service and the organisational culture [ 13 ]. This requires situated knowledge to better understand the emotional work performed by healthcare professionals in specific contexts, such as NHs.

Emotional labour in palliative care and in nursing homes

Emotional labour has been defined as the process through which healthcare professionals suppress or change their feelings to align with organisational rules and guidelines while still conveying to others a sense of being cared for [ 17 ]. Emotional labour involves three strategies: surface acting, deep acting, and displaying genuine emotions [ 18 ]. Surface acting entails suppressing or hiding felt emotions or faking unfelt emotions to alter outwards expression. Deep acting entails a conscious attempt to modify inner feelings and felt emotions to match expected emotions [ 18 , 19 ]. Displaying genuine emotions entails the expression of natural emotions that are congruent with felt feelings without any adjustment [ 18 ]. In this study, we conceptualize emotional labour as the efforts deployed by healthcare professionals to manage their emotions when providing palliative care to NH residents.

Nurse scholars have expanded the concept of emotional labour to nursing, emphasizing the emotional component and the demand associated with caring in nursing [ 19 ]. Nurses perform emotional labour when they have to induce or suppress their feelings to align them with what is expected by their institutions to make patients feel cared for and safe, irrespective of their own actual feelings [ 20 , 21 ]. For example, when nurses are confronted with death but feel unable to facilitate a ‘good death’, they may have complex feelings of guilt and anger but may have to suppress these feelings to continue attending to patients and their relatives [ 22 , 23 ].

Emotions are inherently linked to caring, as they are essential to the development of effective and meaningful relationships with patients, their relatives, and other professionals [ 24 ]. Studies that have explored the emotional labour associated with providing palliative care highlight the complexities of the emotionally demanding experiences healthcare professionals encounter in their practice [ 6 , 23 ]. The cumulative emotional effects of grief and sadness experienced by NH professionals attending to dying individuals require them to deploy significant effort to balance the demands of the healthcare organization, the emotional needs of others, and their own wellbeing [ 23 ].

The limited available literature on emotional labour in NHs illuminates the critical influence that the physical and social environments of NHs have on shaping care providers’ emotional experiences of caring for dying residents [ 7 ]. Caring for ill, disoriented residents with aggressive behaviours as well as dying residents in their last stages of life requires NH professionals to regulate their emotions, often masking their true feelings to prioritize the emotional needs of residents and their families above their own [ 6 , 15 ]. Additionally, ethical and moral concerns that professionals face during end-of-life-care provision, such as preserving residents’ dignity, engaging in end-of-life conversations, respecting end-of-life preferences, life-prolonging treatment or treatment withdrawal, likely play a role in emotional regulation and strategies undertaken by professionals [ 23 , 25 ]. All these may affect healthcare professionals’ capacity to interact effectively with residents and co-workers, nurture their sense of self, and provide optimal care [ 7 , 22 ]. However, it is important to note that the regulation of emotions may also produce positive effects, such as facilitating caring and forming the bonds necessary to foster a home-like NH environment [ 5 , 15 ].

Despite the crucial role that emotions play in providing care in NHs, the emotional labour undertaken by NH care professionals remains an invisible aspect of job requirements [ 15 ]. Notwithstanding the critical role that healthcare professionals play in supporting quality living and quality dying for residents in the French NH context, little is known about how they manage their emotions amidst the complex situations they encounter in their practice or how they continue to provide care without jeopardizing their own emotional wellbeing. This study aimed to help fill this gap in the research by (1) analysing the emotional dimensions of providing palliative care in the NH context as well as the strategies used by healthcare professionals to manage the emotional aspects of caring for dying residents and (2) exploring the effects of emotional labour on NH professionals’ wellbeing. To this end, this study sought to answer the following research questions: (1) How do NH professionals manage the emotional demands of caring for residents requiring palliative care? (2) What effects do emotional demands have on professional-resident interactions and the emotional wellbeing of NH professionals? This study’s findings will inform NH management on strategies and interventions to not only reduce the emotional exhaustion and burnout of healthcare professionals but also improve their resilience and wellbeing at work, ensuring that they are best equipped to provide optimal care to residents and their relatives.

Study overview and design

The findings presented in this manuscript originate from a broader implementation study that evaluated the effectiveness of a timely and integrated palliative care approach in 21 NH facilities in France. The initial study used a mixed-method [ 26 ] approach and was segmented into three phases: pre-implementation, implementation, and post-implementation. The current manuscript reports materials from the pre-implementation phase. A detailed methodological description of the broader study has been reported elsewhere [ 27 ]. The qualitative component of the study follows a multiple case study approach [ 28 ]. Among the aims of the qualitative study were to explore NH professionals’ experiences and quality of life at work and to understand how they navigate the emotional demands associated with providing palliative care to residents.

Participants and settings

This manuscript presents qualitative findings from nine out of the 21 NHs that participated in the broader study. The nine NHs were purposively selected to ensure a balanced sample of three NHs per geographical region (Iles de France, Nouvelle Aquitaine, Provence-Alpes-Côte d’Azur) as well as diversity in terms of ownership status (public, private, and private non-profit).

For this study, we purposively included healthcare professionals from various occupational groups employed on a fixed contract in the selected NHs. Participants had to have a minimum of five months of experience and had experience providing palliative care to residents in the same NH. Casual and non-fixed contractual professionals were excluded from the study.

Data collection

A combination of focus group discussions and individual in-depth interviews was used to collect data. This was to achieve an enhanced understanding of the phenomenon of emotional labour within a NH context by exploring views at individual and social contexts [ 29 ]. Nine focus group discussions were conducted with various healthcare professionals who provide direct care to residents. Each focus group was composed of seven to eleven professionals. Given the purpose of the study which was to explore how different professionals navigate the emotional demands of providing palliative care within a NH context, group composition brought together all professionals involved at varying degrees in providing such a type of care. However, to allow participants to share their experiences freely and to avoid any status distinction or hierarchical influence [ 30 ], professionals in the managerial roles were not included in the focus groups. They were involved in individual interviews.

Prior to data collection, a meeting was organized at each participating NH to introduce the broader interventional study to professionals and invite them to take part in the study. The focus group sessions were held at a predetermined location within the NH at a time convenient to the participants and facility and lasted between 90 and 120 min. Individual in-depth interviews were conducted with the supervisory teams, namely the nurse coordinators and medical coordinators. Each individual interview lasted approximately 45 min and was held at a time and place convenient to participants. We used the interview guide developed by the researchers for the purpose of this study (supplementary material 1). The same interview guide was used for individual and focus group discussions, with slights changes on the phrasing of questions for interviews with the management team. Focus group and individual interview questions inquired about experiences of providing palliative and end-of-life care to residents, the emotional dimensions associated with such a type of care, how professionals navigate those experiences and the perceived consequences on professional wellbeing. Examples of questions included the following: How would you describe your experience of caring for dying residents in NHs? What are the emotional aspects of providing palliative care to residents, and how do you navigate those experiences? The last author conducted most of the individual interviews and some focus group discussions, while a trained research assistant under the supervision of the last author moderated the remaining focus groups. Both hold PhD degrees and have experience in conducting qualitative interviews for health research. All interviews and focus group discussions were conducted from April 2021 to September 2022. The interviews were audio-recorded after the participants granted permission. Regular field notes were written immediately after interviews and focus groups. Data collection continued until we have gained adequate and in-depth understanding [ 31 ] of emotional experiences of providing palliative care in NH.

Data analysis

Thematic content analysis following the analytical approach of Paillé and Mucchielli [ 32 ] guided the analysis. The level of analysis was a NH, with each NH considered a case. After verbatim transcription of all the data from the nine cases, two authors (BU & DL) became acquainted with the data by rereading the transcribed interviews, examining participants’ narratives from each case separately, and then developed a list of codes. From the code list, they created a thematization journal using code subdivision, integration, and hieararchization [ 32 ]. Next, the same two authors grouped related codes from all the cases, with a third team member (BE) resolving any discrepancies between the previous analysts. At the end of this stage, a thematic tree of three themes and eight subthemes was constructed. NVivo 14 software assisted with data management and facilitated the coding process.

To ensure methodological rigor, the authors used the recommended strategies for trustworthiness of qualitative data [ 33 ]. To ensure the reliability of the findings, two analysts completed the coding of transcripts, organized peer debriefing meetings throughout the analysis, and kept a reflexive journal recording all the steps taken and decisions made. A third analyst resolved any disagreements through consensus. To ensure credibility of findings, the authors triangulated methods (individual in-depth interviews and focus groups) and collected participant perspectives from various healthcare professions (nurses, nurse assistants, personal support workers, psychologists, occupational therapists, physiotherapists, medical doctors). The writing of the manuscript followed the “consolidated criteria for reporting qualitative research (COREQ)” [ 34 ] (Supplementary material 2).

Ethical considerations

The French Committee of Protection of Person (CPP) granted ethical approval for this study (Approval number: 2020.09.06 bis_20.07.31.64318). All focus groups and interviews respected the rights of participants to choose to participate in the study through informed consent. To ensure confidentiality and anonymity of the collected data, the reporting used codes instead of names.

Demographic characteristics of participants

All NH professionals who met the inclusion criteria and who were available on the day of the focus group were included in the study. In total, 93 professionals participated, including 79 participants in nine focus groups and 14 participants in individual interviews. Tables  1 and 2 provide the detailed characteristics of the settings and demographic characteristics of participants.

Quotes from individual interviews are followed by an acronym designating the profession of the participant (for example NC for Nurse Coordinator, MedCo for Medical Coordinator) as well as the code number of the NH. Quotes from focus groups are designated by the acronym FG, followed by an acronym for the location of the NH (IDF for Ile de France, NA for Nouvelle Aquitaine, PACA for Provence-Alpes-Côte d’Azur), and the code number of the NH.

Analysis of participants’ narratives revealed three overarching themes related to the emotional dimensions of providing palliative care in NHs: (1) intertwined emotionally rewarding and challenging experiences; (2) multifaceted emotional strategies; and (3) switching between emotional engagement, detachment and exhaustion. Supplementary material 3 illustrates the generation of themes and subthemes with illustrative quotes.

Theme 1. Intertwined emotionally rewarding and challenging experiences

Participants’ narratives revealed two intertwined and simultaneous emotional dimensions of providing palliative care to dying residents in NHs: (1) emotionally rewarding experiences and (2) emotionally challenging experiences. The emotionally rewarding dimension of the experiences was supported by the individual professionals’ intrinsic commitment, devotion, and engagement with older persons under their care and a professional duty to provide them with the “best possible care” they deserve. On the other hand, the NH context as a place of living and of care, with its organisational constraints, rendered the experience of providing palliative care to residents emotionally challenging. Specifically, it hindered the care providers’ ability to facilitate what they perceived to be a “dignified death,” leaving them with feelings of distress, frustration, guilt, and uselessness.

Caring for dying residents: emotionally rewarding experiences

Numerous participants described working in NH as a deliberate professional choice and vocation, stemming from their sense of commitment and engagement to offer dependant older people the care they deserve. The relational dimension associated with caring for NH residents gives meaning to their work and becomes a source of pleasure, satisfaction, and self-worth, as illustrated by the following registered nurse:

“Helping older persons is my passion. I find that there is less invasive care in NHs , and there is a relationship that develops and that I enjoy” (NC , NA , 751).

For the majority of healthcare professionals, this deliberate choice to work in NHs implies that confronting death on an ongoing basis is a professional responsibility. Despite the emotional challenges that come with multiple exposures to death, the participants affirmed their commitment to confront death as part of caring for residents. They held the belief that dying is part of living and that accompanying death is a normal process that comes with caring for the living. The devotion to accompanying residents until the end was perceived by NH professionals as a rewarding experience when they felt they had fulfilled their responsibility of facilitating a dignified death:

“It feels so rewarding to see a resident dying the way they should: with dignity , respect , free of pain and with the best possible comfort. That is what we are here for” (FG , PACA , 935).

Although death is considered an expected life trajectory in NHs, participants recognized that dying older persons are often overlooked as a category of the population requiring adequate palliative care. Their perceived duty to accompany residents until the end demonstrates their commitment to ensuring that dying residents experience comfort and dignity equal to that experienced by individuals dying in settings outside the NH.

Similarly, for some residents, NH professionals are the sole individuals they can bond with at a human level and who can meet their diverse emotional needs. Participants believed their role goes beyond that of care providers. Their drive to go above and beyond and make a difference in the end-of-life trajectory of residents becomes a rewarding experience that provides a sense of pride and self-worth. The NH professionals expressed feeling honoured to be the ones to accompany residents in that ultimate moment, even if it means forgetting themselves:

“Aging comes with many losses and emotional needs: most of them do not want to be here [in the NH]; they feel abandoned by their families , they lose their autonomy over things they used to do , they need to feel cared for. Being there for them through the most important moments of their stay here is very rewarding to us. Basically , in the first place , if they are put here , it is so they do not die alone” (FG , PACA , 931). “It’s truly a phase full of emotions where everything comes out: their past traumas , their anxiety , their worries. We try to hold it together , to forget ourselves a little so we can give them what they deserve…” (FG , NA , 755).

Participants’ accounts bring to the forefront that despite the emotionally laden experience of providing palliative care in an NH context, their commitment and determination to accompany residents in their last living moments make this experience emotionally rewarding.

A living and dying space: emotionally challenging experience

Narratives from healthcare professionals emphasized the context of the NH as being initially designed to serve as a living space. Such a home-like living environment that progressively becomes a place of care and ultimately a place of dying leads to the development of close bonds between NH professionals and residents for an extended period. The emotional and physical closeness formed with residents during their stay transcends the caregiver-resident therapeutic relationship. Healthcare professionals have to navigate these different aspects of the NH and remain professional caregivers, all the while providing a home-like environment. One participant explained:

“I’m here from 8am to 8 pm; we live with them [residents]. I do not call some of them by their names anymore. I call them grandpa , grandma. By living with and caring for them for an extended period , they end up becoming like family. When they die , it breaks our hearts” (FG , PACA , 931).

Boundary-setting issues such as these make the relational dimension of care especially difficult, as NH professionals can become too close to residents, which naturally complicates the transition to end-of-life care. Accompanying death for older persons who they have accompanied for living represented a challenging transition for participants who considered residents like their own relatives, as expressed by the following medical coordinator:

“They’re not just people we care for; we live with them. These are the people for whom we’ve fought for every minute to make life worth living. Professionals look after them almost as if they were their parents. Accompanying someone who is going to die , while you have accompanied them so they can live , is in itself emotionally hard” (MedCo , IDF , 116).

Their strong emotional bond with residents became a challenge for NH professionals when discussing the residents’ end-of-life preferences and the palliative decisions that needed to be made in preparation for the end-of-life care plan. Many shared avoiding these discussions, even when residents prompted them, as they did not feel ready to engage in such emotionally charged conversations.

“Palliative care supposes that we should help them think about their death , but we are unable to do that. We are primarily a place of life , and at the end , there is always death. That’s the complexity of [providing] palliative care in NHs” (NC , NA , 753).

The experience of providing palliative care in NHs was also challenged by structural and organisational constraints such as heavy workloads, a lack of time, and unmanaged pain. As a result, some participants reported that they were unable to provide dying residents with the required relief, which caused NH professionals lingering regret even after the death of the resident.

“That resident who died last week , I would have wished to have been able to stay with them a little longer , hold their hand , put on some music , so that there is a presence , like putting life into the end of life. Unfortunately , I was not able to free myself. And it is difficult to live with such a feeling” (FG , IDF , 116).

Theme 2. Multifaceted emotional strategies

When asked how they cope with the emotional demands of providing palliative care to residents, participants highlighted a diversity of emotional strategies they deployed to protect themselves and to continue fulfilling their caring roles. These ranged from genuinely expressing their emotions to modifying and suppressing their emotions to fit the moment. Modification and suppression of emotions were cited as the most commonly used strategies as opposed to the authentic expression of emotions.

“We shed tears”, “We’ve laughed with” : genuine display of emotions

Participants used expressions such as “We cry with” , “We’ve shed tears” , “We’ve laughed with” to convey the authentic emotional strategies put in place. They genuinely expressed their emotions in situations when they felt a deep connection with residents. Most of these genuine emotional strategies occurred in the moments approaching death or following death.

“All the residents on the floor are like my family. Last time I went to see Mrs. B , when my colleague told me she was dead , I was shocked , truly shocked. It was quick. I cried all my tears. I was so sad. I was unable to continue work because I was crying” (FG , IDF , 111).

In some situations, the NH professional’s personal history prompted the authentic expression of feelings. For example, if a resident’s death mirrored the death of their own loved ones, it made it difficult for them to conceal their true feelings as they usually do. Some referred to it as not being able to pretend to have no feelings.

“I accompanied my dying father in palliative care. Every time there is a death of a resident , it echoes my father’s. We had a resident death not long ago. When I saw him , I completely broke down. I cannot pretend anymore” (FG , IDF , 117).

The authenticity of emotions also manifested through allowing oneself time to grieve the death of a resident. A participant discussed requesting a day off to grieve, but some NHs also provided space for grieving the deceased residents.

“I was truly attached to Mr G. When he died , I took a day off. Everyone else [colleagues] continued living their life… laughing. For me , I could not come to work because I was grieving his death” (FG , IDF , 116).

Other participants also recognized a need for the authentic expression of emotions and requested emotional management support for the team.

“For a resident we’d known for a long time to whom we were very attached , we felt helpless with her end-of-life care , and so we genuinely asked for help. We held a round table… , and we asked for help from an external person” (FG , NA , 751).

While some participants expressed their genuine feelings, many participants across focus groups were in favour of emotions being unnoticed and noted a lack of formalized organisational strategies to deal with emotions. Many NH professionals admitted to frequently modifying their feelings to display emotions that are “acceptable.”

“ You wipe your sadness and put a smile on your face ”: modification of displayed emotions

Some NH professionals described their emotional work as involving frequent switching between sadness and joy to respond to the needs of the moment. Many shared trying to display emotions that were not what they truly felt because the situation at hand required them to convey different, often contradictory emotions; for the participants, this constituted difficult emotional labour.

“On one hand , you’ve got one person who is dying and next to them residents who are living. It is very difficult because you have to go into the room [and] take care of someone who is dying. You are sad because you know you will not see that person again. Nevertheless , the moment you see this person for a few minutes , you have to close the door , wipe your sadness , and put a smile on your face to accompany the next-door resident with a serene face. It is difficult to manage all these emotions at the same time. In one day , you have to give contradictory feelings. You are sad for one person , but at the same time , you have to bring joy to the other resident. You have to show them a different face , and that is not easy” (FG , IDF , 117).

The modification of emotions was compounded in the NH environment for some participants who not only adopted an expression in accordance with what was expected but also tried to set limits and find the appropriate time and space for revealing their true feelings when out of the NH. Participants described this ability to emotionally detach as protective:

“The moment I remove my uniform , I immediately put a different face…. When I reach home , if I have situations that have been painful , I allow myself to be restless and sad; I vent my true feelings…” (FG , PACA , 933).

“ You become numb and move on ”: suppression of feelings

Participants discussed the organisational expectation to suppress emotions in order to continue providing effective care. In a quest to fit into institutional norms, many NH professionals who describe themselves as normally prone to showing their emotions had to learn to suppress them.

“By nature , I am a very sensitive person , but now I keep all my emotions inside of me , and at the end , it becomes difficult to unload. Here , it is not common to open up and show emotions or talk about them. We are expected to keep it to ourselves and move on” (FG , PACA , 933).

Multiple exposures to death and a lack of time and a safe space to grieve deceased residents forced NH professionals to suppress their feelings in an attempt to cope with the distress and to continue providing care to residents.

“How can you display feelings when you have four successive deaths? You become numb and move on. Tomorrow you will have another one. You pretend as if everything is fine but there is a problem…” (FG , IDF , 117).

Theme 3. Switching between emotional engagement, detachment and exhaustion

Participants reported that the emotional labour of providing palliative care in NHs results in both negative and positive consequences. The majority of NH professionals noted that the negative consequences of emotional labour resulted in an inability to provide effective care, and the lack of supportive space to express their feelings caused emotional distress, feelings of guilt, and a sense of failure and powerlessness. Some participants accepted emotions as essential to their caring role and mentioned that they give meaning to their work.

“ Everyone was satisfied ”: enhanced satisfaction and meaning of work

Only a few NHs involved in the study had formalized procedures in place to support the emotional wellbeing of their healthcare professionals. These procedures included formal debriefs, a consultation with a psychologist, and massage therapy. In the majority of the NHs, informal peer-to-peer support was mentioned. In settings where emotions are acknowledged and supported, both professionals and managers reported increased satisfaction when accompanying dying residents.

“We had a resident to whom the team was so emotionally involved. When the end was approaching , emotions were high for both professionals and the resident. We [the supervisory team] requested the intervention of the external palliative care team to introduce a third party in the relationship and gently distance the team without completely disengaging them. At the end , everyone was satisfied , and the resident was properly accompanied. It ended up being one of the memorable end-of-life care for the staff” (NC , NA , 751).

When participants believed that they gave their best up to the end and that the outcome was a peaceful death, they gained a sense of pride and accomplishment. They felt they had attained their mission, which, for many, was one of the reasons they remained working in the NH despite the stressful environment.

“On his passing , the resident was so peaceful , so were relatives. It was a real sense of satisfaction. It is the kind of end-of-life care where you feel you have done the right care and that gives you motivation to stay” (FG , NA , 753).

“ It is heart-breaking ”: a sense of guilt and powerlessness

Suppression or modification of emotions affected the wellbeing of participants as well as the care they provided to residents. Different constraints such as time pressure and competing tasks added to their frustration. Their emotional distress manifested itself as constant feelings of guilt, powerlessness, and a sense of failure for not providing adequate care to residents. One of the most common sources of guilty feelings was when the NH professionals believed that they were unable to offer a peaceful, quality presence during end-of-life moments and that the resident died alone. Dying alone was considered inhumane by participants, as they believed that residents were placed in NHs mainly to ensure they do not die alone.

“Very often you tell yourself , ‘Well , I could have been by her side , tell her a comforting word , play the music she loved , rub her forehands , make sure she had a presence… , but no , she is gone and all alone’. It is not human at all , and you carry this with you for long” (FG , PACA , 933).

Participants also expressed feeling powerless when they saw residents in pain and discomfort, and their inability to provide the required comfort to the dying residents left them with an immense sense of failure and uselessness, which negatively affected their wellbeing and their satisfaction with the work done.

“It breaks your heart to see people suffering like this and little is done about it. It’s heart-breaking to think , ‘We are here to help them , but in fact we’re not even doing that’. We are useless” (FG , NA , 755).

“ You finally give up ”: distancing and exhaustion

To protect themselves from the distress associated with multiple exposures to death and a lack of institutional support, some participants admitted that they banalized death to emotionally distance themselves from dying residents, a strategy that the NH professionals recognized as inadequate.

“We give , we give , one day we can’t take it anymore and we banalize death. We don’t see death anymore. It does not affect us any longer , it becomes a commonplace gesture , mundane. Someone dies today; you attend to the next person who will be gone tomorrow , and so forth and so on. You keep accumulating and one day you explode” (FG , IDF , 117).

Other NH professionals adopted a superficial attitude in an attempt to distance themselves and detach themselves from their true feelings. They chose to involve themselves less in the therapeutic relationship by concentrating more on carrying out instrumental and technical care and less on offering a caring and relational presence.

“I go in [the room] , I give the injection and I get out. Not because I do not want to stay , but because I am thinking of the others. I cannot stay with the one who is dying while I have 70 others who are still alive. I have to look after those who are not dying” (NC , PACA , 935).

Some participants dealt with emotionally challenging situations by refusing to accept the palliative care plan of residents with whom they had strong ties. They would ignore the team’s decisions when it involved withdrawing feeding and restricting movement and instead provide the usual care such as taking blood pressure, providing hydration, and mobilizing residents, irrespective of the futile outcome or the risk of causing more suffering. In this way, they felt more useful towards the residents.

“We had a staff meeting , and they said Mrs X was in end-of-life care…that we should avoid mobilizing her and emphasize comfort care. When I arrived in the room , I did not do anything they said. Rather , I got her up from bed , I washed her , I dressed her , I brought the wheelchair , and I was about to take her out. When the nurse coordinator arrived , she could not understand what I was doing. I was in denial. I could not believe she was dying” (FG , IDF , 116).

Several participants reported feeling emotionally strained, exhausted, and lacking the energy to accomplish their mission. Some of them even resigned or verbalized their intentions to resign from their posts. Their emotional exhaustion reportedly stemmed from an accumulation of frustration, discouragement and a lack of accomplishment, feeling incompetent, and a lack of support, which prompted them to resign rather than form a negative view of the residents and fail to deliver effective care.

“You fight , you try your best to keep going , you get discouraged and finally you give up. That is why I want to do something else. Eventually , I want to take care of people and give them what they deserve. Here , I do not give them what they deserve , which frustrates me , and I accumulate. I resigned. I would rather leave the job to someone who wants to do it the way it is done. As for me , I am going to hold onto something else. I do not want to become a bitter caregiver….” (FG , PACA , 933).

The findings from this study illustrate that providing palliative care to dying residents within the NH context exposes healthcare professionals to intertwined rewarding and exhausting emotional experiences. This emotionally demanding work results in a constant switching between feelings of pride and accomplishment on the one hand and guilt, distress, and grief on the other, and it prompts healthcare professionals to identify and distance themselves from the residents to protect themselves from emotional suffering. These findings lead to greater insights into how NH professionals navigate these emotionally laden situations to meet the needs of the residents and the NH as well as their own needs. Drawing from these important findings, our discussion focuses on three key insights from the study: (1) Caring for dying residents results in both emotionally rewarding and emotionally exhausting experiences, (2) NH professionals have to perform emotional labour to navigate the experiences associated with providing palliative care, and (3) Unrecognized emotional labour undermines the wellbeing of NH professionals.

Caring for dying residents results in both emotionally rewarding and emotionally exhausting experiences

Genuine interest in caring for frail older persons is at the heart of the engagement and commitment demonstrated by the participants in our study. Participants described becoming emotionally attached to residents they care for as “unavoidable and the right thing to do,” especially given the expected “home-like” environment of the NH. In that sense, the affective dimension of working in NH and the internal motivation of the healthcare professionals aligned and helped them navigate the emotional labour of caring for dying residents and added meaning to their work. In line with other studies, the unique characteristics of NH, where care providers and residents engage repeatedly in deep personal and intimate exchanges for an extended time, forged closer and more trusting reciprocal relationships than are typically found within acute care setting nurse-patient interactions [ 35 , 36 ].

A majority of participants recalled the emotionally rewarding experiences associated with caring for frail and dying residents. The NH professionals described accompanying residents as their professional duty and took pride in making their last days as dignified, comfortable, home-like, and respectful as possible. Moreover, accompanying residents in their final moments was considered a moral responsibility by participants. The positive experiences and feelings stemming from close and trusting relationships with residents have been recognized by previous studies as central to the emotional wellbeing of NH professionals [ 15 ]. Direct caregivers for dying residents characterize those particular moments as the rare moments they feel appreciated, noticed, and like they are making a difference in settings where they generally feel unseen [ 36 ]. In particular, our participants expressed positive emotions such as engagement, pride, accomplishment, and self-worth in situations where they felt they had achieved dignity in caring for the dying residents. This is relevant because dignity represents an essential part of caring in NHs and in palliative care [ 37 , 38 ]. These personal characteristics and intrinsic motives constitute the internal resources and resilience attributes that allow healthcare professionals to cope with distressing situations surrounding accompanying death in NHs [ 10 ]. Future interventions and training should aim to reinforce the internal resources of NH professionals with a strong focus on building resilience.

Although participants perceived caring for dying residents as a rewarding experience, when the challenging working conditions within NH hindered them from achieving their moral and professional responsibility, it turned the experience into difficult emotional labour. The current NH working environment fails to provide necessary organisational resources and subsequently creates discrepancies between the ideals held by NH professionals on what constitutes the right comfort care to provide and the current practices. Under severe labour shortages, NHs prioritize technical and task-oriented activities over relational moments [ 39 ]. However, for participants in this study, not being present to hold the hands of the dying resident left them feeling guilty of failing their moral responsibility and their professional duty. Consistent with previous studies, the NH culture was found to prioritize tasks and expect healthcare professionals to be consistently “doing something” for residents versus “being” with residents [ 13 , 40 ]. This dissonance creates the most difficult emotional challenges, moral concerns, and distress for NH professionals [ 25 ]. That perceived inability to facilitate a “good death” due to organisational constraints results in moral distress for NH professionals and complicates their grieving process [ 23 , 41 ]. Echoing this, participants in our study shared how emotionally burdensome it was to constantly feel guilty of devoting less time to the “dying resident” because they were required to attend to the “living residents” instead. NH managers and policymakers should take measures to build a culture that enables healthcare professionals to prioritize the emotional needs of residents alongside their physical care needs, as both are equally important to end-of-life care.

Professionals have to perform emotional labour to navigate the experiences associated with providing palliative care

Participants in the current study used different emotional labour strategies to navigate the rewarding and challenging aspects of caring for dying residents in the NH context. Some adopted distancing strategies, such as focusing on task-based care and mechanical actions as well as avoiding feelings and emotional involvement, while others trivialized death or denied the impending death of residents. This process of strategy switching between engagement and detachment is prevalent among palliative care professionals as a way of coping with emotional demands and preventing grief [ 13 , 23 ].

Numerous participants reported that they tended to modify their feelings by displaying emotions that were different from those they felt. For example, some noted “wiping [their] face and showing a smiley face,” while others suppressed their feelings to “become numb and move on” in an attempt to display composure in the moment and comply with institutional rules. Attempting to modify one’s felt emotions to match displayed emotions is known as deep acting, whereas displaying fake, unfelt emotions and suppressing one’s felt emotions indicates a surface acting strategy [ 18 ]. The emotional strategies used by the participants in our study are similar to those commonly used by healthcare professionals in different care contexts [ 12 ]. Particularly in the NH context, emotional labour is intensified by the long-term therapeutic relationship, as the longer the therapeutic relationship the more complicated the emotional labour [ 7 , 15 ]. Participants in our study shared that the stronger and the closer the bond with the resident, the harder it was to navigate the emotional labour associated with witnessing their suffering and providing them with end-of-life care. Debates persist on the appropriate emotional distance to take when accompanying a resident with whom the healthcare professional has formed a close bond. It is noteworthy, however, that healthcare professionals who try to convey caring while remaining emotionally detached may experience increased emotional dissonance and potentially negative effects [ 23 ]. This phenomenon resonates particularly within the NH care context, where professional boundaries are blurred and difficult to respect [ 6 ].

Some participants in this study identified the importance of safe spaces where they can freely express their emotions without faking and without feeling judged, such as spending time informally with colleagues during breaks or with relatives at home. Researchers classify this as the backstage area of emotional expression, owing to the lack of formal recognition and poor appreciation of emotional labour in practice [ 42 , 43 ]. Given the complexity of emotional labour associated with providing palliative and end-of-life care in NHs, scholars recommend more strengthened, explicit, and structured backstage areas to recognize the emotional needs of healthcare professionals and support their emotional growth and resilience [ 43 ]. Unfortunately, findings from our study reiterate the inadequate support available in the NH context for their mental and emotional wellbeing.

In a few instances, some participants in our study allowed themselves to express naturally felt emotions. The close bond they had formed with residents prompted those who adopted the genuine manifestation of feelings to view the resident’s death as a parallel of their own loved one’s death; hence, they allowed themselves to react accordingly. Some took a leave of absence to process the grief, while others requested formal support as they struggled to come to terms with the death of the resident. Studies have shown that adopting naturally felt emotions as an emotional labour strategy can protect healthcare professionals from burnout, as it allows for authenticity and empathy expressions in care [ 12 ]. Genuine emotions have also been found to support nurses in the provision of compassionate care and to inspire cooperation from less-cooperative residents [ 6 ]. In our study, however, the absence of a formal supportive space within the NH to vent emotions discouraged the genuine expression of feelings. Even in the few NHs where opportunities for emotional sharing existed through support groups and psychologist interventions, the participants were reluctant to take advantage of these opportunities. One possible explanation could be that openly expressing emotions might be seen as a sign of weakness, incompetency, and inability to respect professional boundaries. Yet, organisational studies have shown that when grief and emotional suffering are acknowledged and collectively shared as a team, emotional distress is no longer perceived as an individual weakness but rather a collective suffering that requires collective measures to address. However, this cultural shift is only possible when it is supported by the institution through the provision of time, space, and opportunity to debrief and grieve [ 44 ].

Unrecognized emotional labour undermines the wellbeing of nursing home professionals

This study revealed that the emotions experienced by professionals receive relatively little attention within the NH context. This finding supports other studies that have highlighted the invisible nature of the emotional labour endured by healthcare professionals in end-of-life and palliative care within NHs [ 7 , 15 ]. Current institutional rules reinforced by professional norms such as the self-imposed emotional strategies used by healthcare professionals implicitly discourage the open expression of emotions and position genuine displays of emotion as incompetence [ 6 ]. Considering the expression of emotions as weak and a sign of a problem to be addressed leads to emotional labour being unrecognized, professionally undervalued, and even discriminated against [ 24 ]. This is deeply problematic, as unrecognized emotional labour can lead to personal, professional, and organisational negative outcomes.

The effect of emotional labour on a healthcare professional varies depending on the frequency, intensity, diversity, and length of the needed emotional displays as well as the degree of emotional dissonance between the emotions experienced and those anticipated [ 14 ]. Evidence demonstrates that a constant mismatch between felt feelings and displayed emotions leads to emotional dissonance, an internal state of conflict that can cause healthcare professionals to experience difficulty in patient interactions, high levels of stress and burnout [ 12 ], emotional “estrangement” (p.443) [ 13 ], and increased intention to leave [ 45 ]. Consistent with these studies, participants in our study felt drained and worn out by the emotional efforts associated with caring for the dying residents. They experienced guilt and feelings of powerlessness and failure, and a significant number expressed their intention to leave the NH.

At the organisational level, studies have demonstrated that poor patient outcomes and poor quality of care, including missing care opportunities and mistreating residents, are potential negative outcomes of emotional exhaustion and unrecognized emotional labour, as well as lower levels of staff recruitment and retention [ 46 ]. In contrast, emerging evidence suggests that when institutional expectations allow and support authentic emotional expression, positive effects can result for healthcare professionals, care recipients, and the healthcare system [ 12 , 14 ].

Implications for policy, practice, and research

The findings from this study expand our understanding of the complex emotional demands associated with caring for dying residents in NHs. Caring for frail older persons requires extensive time, effort, and mental and physical energy, and it involves the interplay of physical and emotional tasks and skills [ 13 ]. These findings represent a valuable contribution to the NH care system, a system that has been predominated by instrumental-focused care. The data highlights the need for a paradigm shift toward valuing the emotional labour involved in accompanying life and death in contexts that are not palliative-care specialised. Upholding quality care alongside the wellbeing of healthcare professionals requires an organisational culture that does not separate instrumental acts from the emotional labour at the very heart of the caring profession. Instead, it requires organisational changes that result in emotional support seen as a collective routine practice that strengthens the team rather than as an individual responsibility and weakness. This will allow NH professionals to regularly share their feelings and emotions, leading to emotional openness and acceptance [ 40 ].

Regular in-service training initiatives should be put in place in NHs to equip healthcare professionals with effective emotional management skills. In particular, the nursing assistants and personal support workers in our study appeared to be most affected by the negative impact of emotional labour. These categories of professional groups require tailored training to help bridge their skill gap. Capacity building approaches such as critical companionship have been proven to equip healthcare professionals with skills on the effective use of emotions in therapeutic relationships and to allow them to reflect on the use of self in caring [ 19 ]. As a lack of institutional support and peer support discourages emotional expression, NH settings should reinforce work environments in which leadership, supervisor, and co-worker support are an integral part of routine practices.

Structural deficiencies such as inadequate staffing, heavy workloads, and competing tasks leave NH healthcare professionals with inadequate time to provide optimal care. This underpins most of the challenges healthcare professionals experience in the NH context and is a primary factor in the emotional burden they experience when they fail to provide quality palliative care to dying residents. There is a need to adapt resource allocation to the complexity of providing palliative care within NHs. Further studies are needed to design interventions that support emotional regulation while increasing the resilience and emotional intelligence of healthcare professionals in NHs.

A strength of our study was the use of individual and focus groups interviews, which enabled a comprehensive exploration of individual and group views on emotional labour of NH professionals. Including professionals involved in direct care and leaders, i.e. nurse and medical coordinators, enabled to capture a diverse set of experiences and perspectives across professional categories and roles. This study did not intend to establish the levels of influence of factors such as professional category, years of work experience, level of interaction with residents or settings characteristics on emotional labour and strategies used. This may constitute the focus of future research.

This study brought to the forefront the complex emotional labour performed by NH professionals while caring for residents requiring palliative care. The results demonstrated that emotions are an undeniable part of caring for frail and dying older persons in the context of a home-like environment; however, current NH culture discourages genuine emotional sharing and emphasizes emotional suppression. Unrecognized emotions undermine the wellbeing of healthcare professionals, leading to negative individual and organisational outcomes. Understanding and acknowledging the emotional labour of NH professionals is critical to supporting their wellbeing, resilience, and retention, and it ultimately may improve the quality of care for dying residents. The stigma surrounding the emotional labour of caring can be broken by decision makers who design healthy workplace environments that celebrate emotional transparency as a strength as well as by each and every healthcare worker who bravely displays their genuine emotions in hopes to shape a new culture that fully acknowledges their humanity alongside their professional skills.

Data availability

The datasets used in this study are available on a reasonable request from the corresponding author.

Honinx E, Van den Block L, Piers R, Onwuteaka-Philipsen B, Payne S, Szczerbińska K, et al. Large differences in the organization of palliative care in nursing homes in six European countries: findings from the PACE cross-sectional study. BMC Palliat care. 2021;20:1–12.

Article   Google Scholar  

Carpenter JG, Lam K, Ritter AZ, Ersek M. A systematic review of nursing home palliative care interventions: characteristics and outcomes. J Am Med Dir Assoc. 2020;21(5):583–96. e2.

Article   PubMed   PubMed Central   Google Scholar  

Reitinger E, Froggatt K, Brazil K, Heimerl K, Hockley J, Kunz R, et al. Palliative care in long-term care settings for older people: findings from an EAPC taskforce. Eur J Palliat Care. 2013;20(5):251–3.

Google Scholar  

Froggatt K, Payne S, Morbey H, Edwards M, Finne-Soveri H, Gambassi G, et al. Palliative care development in European care homes and nursing homes: application of a typology of implementation. J Am Med Dir Assoc. 2017;18(6):550. e7-. e14.

Article   PubMed Central   Google Scholar  

Sussman T, Kaasalainen S, Mintzberg S, Sinclair S, Young L, Ploeg J, et al. Broadening end-of-life comfort to improve palliative care practices in long term care. Can J Aging/La Revue Canadienne Du Vieillissement. 2017;36(3):306–17.

Yan Z. I tried to control my emotions: nursing home care workers’ experiences of emotional labor in China. J Cross-Cult Gerontol. 2022;37(1):1–22.

Giesbrecht M, Stajduhar KI, Cloutier D, Dujela C. We are to be like machines… fill the bed before it gets cold: exploring the emotional geographies of healthcare providers caring for dying residents in long-term care facilities. Soc Sci Med. 2021;272:113749.

Article   PubMed   Google Scholar  

Beyrie A, Douguet F, Mino J-C. Accompagner la fin de vie en maison de retraite: spécificités et limites d’un travail inestimable. ethnographiques org. 2017. (35-Vieillir en institution, vieillesses institutionnalisées. Nouvelles populations, nouveaux lieux, nouvelles pratiques).

Curtin M, Richards HL, Fortune DG. Resilience among health care workers while working during a pandemic: a systematic review and meta synthesis of qualitative studies. Clin Psychol Rev. 2022;95:102173.

Chang SO, Kim EY. The resilience of nursing staffs in nursing homes: concept development applying a hybrid model. BMC Nurs. 2022;21(1):129.

Yeh S-CJ, Chen S-HS, Yuan K-S, Chou W, Wan TT. Emotional labor in health care: the moderating roles of personality and the mediating role of sleep on job performance and satisfaction. Front Psychol. 2020;11:574898.

Kim JS. Emotional labor strategies, stress, and burnout among hospital nurses: a path analysis. J Nurs Scholarsh. 2020;52(1):105–12.

Hanlon N. Professional caring in affective services: the ambivalence of emotional nurture in practice. Eur J Social Work. 2023;26(3):441–53.

Xu Y-w. Emotional labor and job satisfaction among nurses: the mediating effect of nurse–patient relationship. Front Psychol. 2023;14:1094358.

Franzosa E, Tsui EK, Baron S. Who’s caring for us? Understanding and addressing the effects of emotional labor on home health aides’ well-being. Gerontologist. 2019;59(6):1055–64.

Wen J, Huang SS, Hou P. Emotional intelligence, emotional labor, perceived organizational support, and job satisfaction: a moderated mediation model. Int J Hospitality Manage. 2019;81:120–30.

Hochschild AR. Travail émotionnel, règles de sentiments et structure sociale 1. Travailler. 2003;1:019–49.

Diefendorff JM, Croyle MH, Gosserand RH. The dimensionality and antecedents of emotional labor strategies. J Vocat Behav. 2005;66(2):339–57.

Smith P, Gray B. Emotional labour of nursing revisited: caring and learning 2000. Nurse Educ Pract. 2001;1(1):42–9.

Badolamenti S, Sili A, Caruso R, FidaFida R. What do we know about emotional labour in nursing? A narrative review. Br J Nurs. 2017;26(1):48–55.

Gray B, Smith P. Emotional labour and the clinical settings of nursing care: the perspectives of nurses in East London. Nurse Educ Pract. 2009;9(4):253–61.

Ingebretsen LP, Sagbakken M. Hospice nurses’ emotional challenges in their encounters with the dying. Int J Qualitative Stud Health well-being. 2016;11(1):31170.

Funk LM, Peters S, Roger KS. The emotional labor of personal grief in palliative care: balancing caring and professional identities. Qual Health Res. 2017;27(14):2211–21.

Fernández-Basanta S, Lois‐Sandá L, Movilla‐Fernández MJ. The link between task‐focused care and care beyond technique: a meta‐ethnography about the emotional labour in nursing care. J Clin Nurs. 2023;32(13–14):3130–43.

Wiersma E, Marcella J, McAnulty J, Kelley ML. That just breaks my heart’: moral concerns of direct care workers providing palliative care in LTC homes. Can J Aging/La Revue Canadienne Du Vieillissement. 2019;38(3):268–80.

Guével M-R, Pommier J. Recherche par les méthodes mixtes en santé publique: Enjeux Et illustration. Santé Publique. 2012;24(1):23–38.

Bagaragaza E, Colombet I, Perineau M, Aegerter P, Guirimand F. Assessing the implementation and effectiveness of early integrated palliative care in long-term care facilities in France: an interventional mixed-methods study protocol. BMC Palliat care. 2023;22(1):35.

Yin RK. Case study research and applications. Sage Thousand Oaks, CA; 2018.

Lambert SD, Loiselle CG. Combining individual interviews and focus groups to enhance data richness. J Adv Nurs. 2008;62(2):228–37.

Patton MQ. Qualitative research & evaluation methods. 3rd ed. Thousand Oaks, CA: Sage; 2002.

Braun V, Clarke V. To saturate or not to saturate? Questioning data saturation as a useful concept for thematic analysis and sample-size rationales. Qualitative Res Sport Exerc Health. 2021;13(2):201–16.

Paillé P, Mucchielli A. L’analyse qualitative en sciences humaines et sociales-5e éd. Armand Colin; 2021.

Maher C, Hadfield M, Hutchings M, De Eyto A. Ensuring rigor in qualitative data analysis: a design research approach to coding combining NVivo with traditional material methods. Int J Qualitative Methods. 2018;17(1):1609406918786362.

Tong A, Sainsbury P, Craig J. Consolidated criteria for reporting qualitative research (COREQ): a 32-item checklist for interviews and focus groups. Int J Qual Health Care. 2007;19(6):349–57.

Rodriquez J, editor. Editor it’sa dignity thing: nursing Home Care Workers’ Use of emotions 1. Sociological forum. Wiley Online Library; 2011.

Cagle JG, Unroe KT, Bunting M, Bernard BL, Miller SC. Caring for dying patients in the nursing home: voices from frontline nursing home staff. J Pain Symptom Manag. 2017;53(2):198–207.

Holmberg B, Godskesen T. Dignity in Care at the End of Life in a Nursing Home: an Ethnographic Study. 2022.

Östlund U, Blomberg K, Söderman A, Werkander Harstäde C. How to conserve dignity in palliative care: suggestions from older patients, significant others, and healthcare professionals in Swedish municipal care. BMC Palliat care. 2019;18:1–12.

Herr M, Raharimanana S, Bagaragaza E, Aegerter P, Sipos I, Fabre C, et al. Evaluation de la culture de sécurité en etablissement d’Hébergement pour Personnes Agées Dépendantes (EHPAD): adaptation française du questionnaire nursing Home Survey on Patient Safety Culture. Can J Aging/La Revue Canadienne Du Vieillissement. 2017;36(4):453–62.

Funk LM, Peters S, Roger KS. Caring about dying persons and their families: interpretation, practice and emotional labour. Health Soc Care Commun. 2018;26(4):519–26.

Young A, Froggatt K, Brearley SG. Powerlessness’ or ‘doing the right thing’–Moral distress among nursing home staff caring for residents at the end of life: an interpretive descriptive study. Palliat Med. 2017;31(9):853–60.

Bolton S. Changing faces: nurses as emotional jugglers. Sociol Health Illn. 2001;23(1):85–100.

Slootman M, Mudde AL, The AM. Frontstage nursing and backstage growth: the emotional labour of student nurses in Dutch nursing homes. Nurs Inq. 2023;30(4):e12570.

Simard J. Autour de la souffrance soignante en soins palliatifs: entre idéal de Soin et juste distance 1. Frontières. 2014;26(1).

Bartram T, Casimir G, Djurkovic N, Leggat SG, Stanton P. Do perceived high performance work systems influence the relationship between emotional labour, burnout and intention to leave? A study of Australian nurses. J Adv Nurs. 2012;68(7):1567–78.

Eltaybani S, Yamamoto-Mitani N, Ninomiya A, Igarashi A. The association between nurses’ burnout and objective care quality indicators: a cross-sectional survey in long-term care wards. BMC Nurs. 2021;20:1–10.

Download references

Acknowledgements

The authors of this article sincerely thank the funders of this study cited above and the scientific committee members for their valuable support. We acknowledge the contributions of all members of the Padi-Palli team. We are also grateful to the nursing homes and professionals who participated in the study. The authors thank Professor Margaret Fitch for her valuable insights into the manuscript.

This study was supported by the French Ministry of Health and Solidarity through a call for projects PREPS (Healthcare System Performance Research Program): grant number PREPS 19–0066, by the Association des Dames du Calvaire (ADC) and by the Regional Health Agencies (Agence Régionale de Santé) of Ile de France (ARS IDF) and Provence-Alpes-Côte d’Azur (ARS PACA). The funders had no role or responsibilities in the study design, data collection, data management, analysis and interpretation, or publication of this manuscript.

Author information

Authors and affiliations.

Département Recherche Enseignement Formation, Etablissements Jeanne Garnier, 106 avenue Émile Zola, Paris, 75015, France

Benoite Umubyeyi, Danièle Leboul & Emmanuel Bagaragaza

You can also search for this author in PubMed   Google Scholar

Contributions

EB conceptualized and designed the study, collected and analysed the data, and revised the manuscript. BU analysed the data and drafted and revised the manuscript. DL analysed the data and revised the manuscript. All authors have reviewed and approved the final manuscript.

Corresponding author

Correspondence to Emmanuel Bagaragaza .

Ethics declarations

Ethics approval and consent to participate.

Research ethics approval for this study was granted by the French Committee of Protection of Person, approval number 2020.09.06 bis 20.07.31.64318. The study is registered in the National Study Database as ID-RCB 2020-A01832-37. The use of databases and data processing were implemented in accordance with French law (“Informatique et Libertés” dated January 6, 1978 and amended June 20, 2018) and European regulations (General Data Protection Regulation - GDPR dated April 27, 2016). All participants provided their informed consent in writing before their inclusion in the study. Confidentiality was ensured using codes and pseudonyms.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note.

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

Supplementary Information

Supplementary material 1., supplementary material 2., supplementary material 3., rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, 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 you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. 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-nc-nd/4.0/ .

Reprints and permissions

About this article

Cite this article.

Umubyeyi, B., Leboul, D. & Bagaragaza, E. “ You close the door , wipe your sadness and put on a smiling face ”: a qualitative study of the emotional labour of healthcare professionals providing palliative care in nursing homes in France. BMC Health Serv Res 24 , 1070 (2024). https://doi.org/10.1186/s12913-024-11550-7

Download citation

Received : 28 May 2024

Accepted : 05 September 2024

Published : 16 September 2024

DOI : https://doi.org/10.1186/s12913-024-11550-7

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Nursing home
  • Emotional labour
  • Staff wellbeing
  • Palliative care
  • End-of-life care
  • Emotional labour strategies
  • Healthy environment
  • Healthcare professionals

BMC Health Services Research

ISSN: 1472-6963

qualitative meaning of research

IMAGES

  1. 11 Types of Qualitative Research

    qualitative meaning of research

  2. Quantitative Methods Examples

    qualitative meaning of research

  3. 16 Qualitative Methods Examples (2024)

    qualitative meaning of research

  4. 16 Qualitative Methods Examples (2024)

    qualitative meaning of research

  5. Qualitative Versus Quantitative Research

    qualitative meaning of research

  6. What is Qualitative Research? Definition, Types, Examples, Methods, and Best Practices

    qualitative meaning of research

VIDEO

  1. 10 Difference Between Qualitative and Quantitative Research (With Table)

  2. What is Qualitative Research

  3. Research process

  4. Class 11 Statistics|| Ch # 01 Defines||Variables|| Types of Variables|| Functions of Statistics 📊

  5. Difference between Qualitative & Quantitative Research

  6. Significance of Qualitative Research Methods 2024

COMMENTS

  1. What Is Qualitative Research?

    Qualitative research involves collecting and analyzing non-numerical data (e.g., text, video, or audio) to understand concepts, opinions, or experiences.

  2. Qualitative Research

    Qualitative research aims to uncover the meaning and significance of social phenomena, and it typically involves a more flexible and iterative approach to data collection and analysis compared to quantitative research. Qualitative research is often used in fields such as sociology, anthropology, psychology, and education.

  3. Definition

    Definition Qualitative research is the naturalistic study of social meanings and processes, using interviews, observations, and the analysis of texts and images. In contrast to quantitative researchers, whose statistical methods enable broad generalizations about populations (for example, comparisons of the percentages of U.S. demographic groups who vote in particular ways), qualitative ...

  4. What is Qualitative in Qualitative Research

    Qualitative research is multimethod in focus, involving an interpretative, naturalistic approach to its subject matter. This means that qualitative researchers study things in their natural settings, attempting to make sense of, or interpret, phenomena in terms of the meanings people bring to them.

  5. What Is Qualitative Research? An Overview and Guidelines

    This guide explains the focus, rigor, and relevance of qualitative research, highlighting its role in dissecting complex social phenomena and providing in-depth, human-centered insights. The guide also examines the rationale for employing qualitative methods, underscoring their critical importance. An exploration of the methodology's ...

  6. Qualitative research

    Qualitative research is a type of research that aims to gather and analyse non-numerical (descriptive) data in order to gain an understanding of individuals' social reality, including understanding their attitudes, beliefs, and motivation.

  7. What is Qualitative Research? Definition, Types, Examples ...

    Qualitative research is defined as an exploratory method that aims to understand complex phenomena, often within their natural settings, by examining subjective experiences, beliefs, attitudes, and behaviors. Learn more about qualitative research methods, types, examples and best practices.

  8. What Is Qualitative Research?

    Qualitative research is used to understand how people experience the world. While there are many approaches to qualitative research, they tend to be flexible and focus on retaining rich meaning when interpreting data. Common approaches include grounded theory, ethnography, action research, phenomenological research, and narrative research.

  9. The SAGE Encyclopedia of Qualitative Research Methods

    Qualitative research is designed to explore the human elements of a given topic, while specific qualitative methods examine how individuals see and experience the world. Qualitative approaches are typically used to explore new phenomena and to capture individuals' thoughts, feelings, or interpretations of meaning and process.

  10. Qualitative Study

    Qualitative research gathers participants' experiences, perceptions, and behavior. It answers the hows and whys instead of how many or how much. It could be structured as a standalone study, purely relying on qualitative data, or part of mixed-methods research that combines qualitative and quantitative data.

  11. Planning Qualitative Research: Design and Decision Making for New

    For students and novice researchers, the choice of qualitative approach and subsequent alignment among problems, research questions, data collection, and data a...

  12. Qualitative Research: An Overview

    Qualitative research is one of the most commonly used types of research and methodology in the social sciences. Unfortunately, qualitative research is commonly misunderstood. In this chapter, we describe and explain the misconceptions surrounding qualitative research...

  13. What is Qualitative Research? Methods, Types, Approaches and Examples

    Qualitative research is the process of collecting, analyzing, and interpreting non-numerical data. The findings of qualitative research are expressed in words and help in understanding individuals' subjective perceptions about an event, condition, or subject. This type of research is exploratory and is used to generate hypotheses or theories ...

  14. The Oxford Handbook of Qualitative Research

    Abstract. The Oxford Handbook of Qualitative Research, second edition, presents a comprehensive retrospective and prospective review of the field of qualitative research. Original, accessible chapters written by interdisciplinary leaders in the field make this a critical reference work. Filled with robust examples from real-world research ...

  15. How to use and assess qualitative research methods

    This paper aims to provide an overview of the use and assessment of qualitative research methods in the health sciences. Qualitative research can be defined as the study of the nature of phenomena and is especially appropriate for answering questions ...

  16. 4.2 Definitions and Characteristics of Qualitative Research

    4.2 Definitions and Characteristics of Qualitative Research Qualitative research aims to uncover the meaning and understanding of phenomena that cannot be broken down into measurable elements. It is based on naturalistic, interpretative and humanistic notions. 5 This research method seeks to discover, explore, identify or describe subjective human experiences using non-statistical methods and ...

  17. Qualitative Research Definition and Methods

    Qualitative research is a type of social science research that collects and works with non-numerical data and that seeks to interpret meaning from these data that help understand social life through the study of targeted populations or places. People often frame it in opposition to quantitative research, which uses numerical data to identify ...

  18. What is qualitative research?

    Traditionally, in the health sciences, qualitative research has been defined in opposition to quantitative research. A stereotypical view of qualitative research is that it is defined by its data generation methods, such as interviews and observations.

  19. Qualitative vs Quantitative Research: What's the Difference?

    The main difference between quantitative and qualitative research is the type of data they collect and analyze. Quantitative data is information about quantities, and therefore numbers, and qualitative data is descriptive, and regards phenomenon which can be observed but not measured, such as language. Quantitative research collects numerical ...

  20. What is Qualitative Research? Methods and Examples

    Qualitative research seeks to understand people's experiences and perspectives by studying social organizations and human behavior. Data in qualitative studies focuses on people's beliefs and emotional responses. Qualitative data is especially helpful when a company wants to know how customers feel about a product or service, such as in ...

  21. Conceptualization in qualitative research

    Qualitative research questions often ask about lived experience, personal experience, understanding, meaning, and stories. Qualitative research questions may be more general and less specific. Qualitative research questions may also contain only one variable, rather than asking about relationships between multiple variables.

  22. Qualitative Study

    Qualitative research is a type of research that explores and provides deeper insights into real-world problems. Instead of collecting numerical data points or intervening or introducing treatments just like in quantitative research, qualitative research helps generate hypothenar to further investigate and understand quantitative data.

  23. Qualitative vs. Quantitative Research

    When collecting and analyzing data, quantitative research deals with numbers and statistics, while qualitative research deals with words and meanings.

  24. Challenges of conducting qualitative social work research with older

    The research findings suggest that researchers may encounter various challenges when conducting qualitative social work research with older adults, despite the fact that a prior educational background in social work provides an advantage. It is anticipated that the visibility of qualitative research experiences with older individuals will ...

  25. Gaps in communication theory paradigms when conducting implementation

    Background Communication is considered an inherent element of nearly every implementation strategy. Often it is seen as a means for imparting new information between stakeholders, representing a Transaction orientation to communication. From a Process orientation, communication is more than information-exchange and is acknowledged as being shaped by (and shaping) the individuals involved and ...

  26. What is Qualitative in Qualitative Research

    Qualitative research is multimethod in focus, involving an interpretative, naturalistic approach to its subject matter. This means that qualitative researchers study things in their natural settings, attempting to make sense of, or interpret, phenomena in terms of the meanings people bring to them.

  27. Social Determinants Influencing the Non-Adoption of Norms Favorable to

    A qualitative exploratory descriptive design guided data collection and analysis 27,28 in accordance with the COnsolidated criteria for REporting Qualitative research (COREQ) guideline. 29 This approach was chosen because of its interpretative, spontaneous, and natural way of approaching, questioning, and understanding realities. 30

  28. (PDF) Perceptions of burnout among public sector ...

    This is an exploratory qualitative study employing in-depth structured interviews with 24 par- ticipants between April 1, 2022 and July 31, 2022. This study was designed through a series of

  29. "You close the door, wipe your sadness and put on a smiling face": a

    Palliative care provided to frail and dying older persons in nursing homes results in intense emotions for residents and their relatives as well as for healthcare professionals. In France, scant attention has been given to how nursing home professionals manage their emotions when providing palliative care. This study analysed the emotional demands of providing palliative care in the nursing ...

  30. Surface and Depth in Sensory Ethnography: Casting Bodied Experience in

    Arts-based methods developed concurrently in qualitative research and art therapy circles in the 1970s (Leavy, 2015). That similar methods are shared across therapeutic and research domains raises critical questions about how to differentiate therapeutic practice from research. Sensory and imaginative ethnographic investigations that mobilize ...