Logo for M Libraries Publishing

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

7.3 Quasi-Experimental Research

Learning objectives.

  • Explain what quasi-experimental research is and distinguish it clearly from both experimental and correlational research.
  • Describe three different types of quasi-experimental research designs (nonequivalent groups, pretest-posttest, and interrupted time series) and identify examples of each one.

The prefix quasi means “resembling.” Thus quasi-experimental research is research that resembles experimental research but is not true experimental research. Although the independent variable is manipulated, participants are not randomly assigned to conditions or orders of conditions (Cook & Campbell, 1979). Because the independent variable is manipulated before the dependent variable is measured, quasi-experimental research eliminates the directionality problem. But because participants are not randomly assigned—making it likely that there are other differences between conditions—quasi-experimental research does not eliminate the problem of confounding variables. In terms of internal validity, therefore, quasi-experiments are generally somewhere between correlational studies and true experiments.

Quasi-experiments are most likely to be conducted in field settings in which random assignment is difficult or impossible. They are often conducted to evaluate the effectiveness of a treatment—perhaps a type of psychotherapy or an educational intervention. There are many different kinds of quasi-experiments, but we will discuss just a few of the most common ones here.

Nonequivalent Groups Design

Recall that when participants in a between-subjects experiment are randomly assigned to conditions, the resulting groups are likely to be quite similar. In fact, researchers consider them to be equivalent. When participants are not randomly assigned to conditions, however, the resulting groups are likely to be dissimilar in some ways. For this reason, researchers consider them to be nonequivalent. A nonequivalent groups design , then, is a between-subjects design in which participants have not been randomly assigned to conditions.

Imagine, for example, a researcher who wants to evaluate a new method of teaching fractions to third graders. One way would be to conduct a study with a treatment group consisting of one class of third-grade students and a control group consisting of another class of third-grade students. This would be a nonequivalent groups design because the students are not randomly assigned to classes by the researcher, which means there could be important differences between them. For example, the parents of higher achieving or more motivated students might have been more likely to request that their children be assigned to Ms. Williams’s class. Or the principal might have assigned the “troublemakers” to Mr. Jones’s class because he is a stronger disciplinarian. Of course, the teachers’ styles, and even the classroom environments, might be very different and might cause different levels of achievement or motivation among the students. If at the end of the study there was a difference in the two classes’ knowledge of fractions, it might have been caused by the difference between the teaching methods—but it might have been caused by any of these confounding variables.

Of course, researchers using a nonequivalent groups design can take steps to ensure that their groups are as similar as possible. In the present example, the researcher could try to select two classes at the same school, where the students in the two classes have similar scores on a standardized math test and the teachers are the same sex, are close in age, and have similar teaching styles. Taking such steps would increase the internal validity of the study because it would eliminate some of the most important confounding variables. But without true random assignment of the students to conditions, there remains the possibility of other important confounding variables that the researcher was not able to control.

Pretest-Posttest Design

In a pretest-posttest design , the dependent variable is measured once before the treatment is implemented and once after it is implemented. Imagine, for example, a researcher who is interested in the effectiveness of an antidrug education program on elementary school students’ attitudes toward illegal drugs. The researcher could measure the attitudes of students at a particular elementary school during one week, implement the antidrug program during the next week, and finally, measure their attitudes again the following week. The pretest-posttest design is much like a within-subjects experiment in which each participant is tested first under the control condition and then under the treatment condition. It is unlike a within-subjects experiment, however, in that the order of conditions is not counterbalanced because it typically is not possible for a participant to be tested in the treatment condition first and then in an “untreated” control condition.

If the average posttest score is better than the average pretest score, then it makes sense to conclude that the treatment might be responsible for the improvement. Unfortunately, one often cannot conclude this with a high degree of certainty because there may be other explanations for why the posttest scores are better. One category of alternative explanations goes under the name of history . Other things might have happened between the pretest and the posttest. Perhaps an antidrug program aired on television and many of the students watched it, or perhaps a celebrity died of a drug overdose and many of the students heard about it. Another category of alternative explanations goes under the name of maturation . Participants might have changed between the pretest and the posttest in ways that they were going to anyway because they are growing and learning. If it were a yearlong program, participants might become less impulsive or better reasoners and this might be responsible for the change.

Another alternative explanation for a change in the dependent variable in a pretest-posttest design is regression to the mean . This refers to the statistical fact that an individual who scores extremely on a variable on one occasion will tend to score less extremely on the next occasion. For example, a bowler with a long-term average of 150 who suddenly bowls a 220 will almost certainly score lower in the next game. Her score will “regress” toward her mean score of 150. Regression to the mean can be a problem when participants are selected for further study because of their extreme scores. Imagine, for example, that only students who scored especially low on a test of fractions are given a special training program and then retested. Regression to the mean all but guarantees that their scores will be higher even if the training program has no effect. A closely related concept—and an extremely important one in psychological research—is spontaneous remission . This is the tendency for many medical and psychological problems to improve over time without any form of treatment. The common cold is a good example. If one were to measure symptom severity in 100 common cold sufferers today, give them a bowl of chicken soup every day, and then measure their symptom severity again in a week, they would probably be much improved. This does not mean that the chicken soup was responsible for the improvement, however, because they would have been much improved without any treatment at all. The same is true of many psychological problems. A group of severely depressed people today is likely to be less depressed on average in 6 months. In reviewing the results of several studies of treatments for depression, researchers Michael Posternak and Ivan Miller found that participants in waitlist control conditions improved an average of 10 to 15% before they received any treatment at all (Posternak & Miller, 2001). Thus one must generally be very cautious about inferring causality from pretest-posttest designs.

Does Psychotherapy Work?

Early studies on the effectiveness of psychotherapy tended to use pretest-posttest designs. In a classic 1952 article, researcher Hans Eysenck summarized the results of 24 such studies showing that about two thirds of patients improved between the pretest and the posttest (Eysenck, 1952). But Eysenck also compared these results with archival data from state hospital and insurance company records showing that similar patients recovered at about the same rate without receiving psychotherapy. This suggested to Eysenck that the improvement that patients showed in the pretest-posttest studies might be no more than spontaneous remission. Note that Eysenck did not conclude that psychotherapy was ineffective. He merely concluded that there was no evidence that it was, and he wrote of “the necessity of properly planned and executed experimental studies into this important field” (p. 323). You can read the entire article here:

http://psychclassics.yorku.ca/Eysenck/psychotherapy.htm

Fortunately, many other researchers took up Eysenck’s challenge, and by 1980 hundreds of experiments had been conducted in which participants were randomly assigned to treatment and control conditions, and the results were summarized in a classic book by Mary Lee Smith, Gene Glass, and Thomas Miller (Smith, Glass, & Miller, 1980). They found that overall psychotherapy was quite effective, with about 80% of treatment participants improving more than the average control participant. Subsequent research has focused more on the conditions under which different types of psychotherapy are more or less effective.

Han Eysenck

In a classic 1952 article, researcher Hans Eysenck pointed out the shortcomings of the simple pretest-posttest design for evaluating the effectiveness of psychotherapy.

Wikimedia Commons – CC BY-SA 3.0.

Interrupted Time Series Design

A variant of the pretest-posttest design is the interrupted time-series design . A time series is a set of measurements taken at intervals over a period of time. For example, a manufacturing company might measure its workers’ productivity each week for a year. In an interrupted time series-design, a time series like this is “interrupted” by a treatment. In one classic example, the treatment was the reduction of the work shifts in a factory from 10 hours to 8 hours (Cook & Campbell, 1979). Because productivity increased rather quickly after the shortening of the work shifts, and because it remained elevated for many months afterward, the researcher concluded that the shortening of the shifts caused the increase in productivity. Notice that the interrupted time-series design is like a pretest-posttest design in that it includes measurements of the dependent variable both before and after the treatment. It is unlike the pretest-posttest design, however, in that it includes multiple pretest and posttest measurements.

Figure 7.5 “A Hypothetical Interrupted Time-Series Design” shows data from a hypothetical interrupted time-series study. The dependent variable is the number of student absences per week in a research methods course. The treatment is that the instructor begins publicly taking attendance each day so that students know that the instructor is aware of who is present and who is absent. The top panel of Figure 7.5 “A Hypothetical Interrupted Time-Series Design” shows how the data might look if this treatment worked. There is a consistently high number of absences before the treatment, and there is an immediate and sustained drop in absences after the treatment. The bottom panel of Figure 7.5 “A Hypothetical Interrupted Time-Series Design” shows how the data might look if this treatment did not work. On average, the number of absences after the treatment is about the same as the number before. This figure also illustrates an advantage of the interrupted time-series design over a simpler pretest-posttest design. If there had been only one measurement of absences before the treatment at Week 7 and one afterward at Week 8, then it would have looked as though the treatment were responsible for the reduction. The multiple measurements both before and after the treatment suggest that the reduction between Weeks 7 and 8 is nothing more than normal week-to-week variation.

Figure 7.5 A Hypothetical Interrupted Time-Series Design

A Hypothetical Interrupted Time-Series Design - The top panel shows data that suggest that the treatment caused a reduction in absences. The bottom panel shows data that suggest that it did not

The top panel shows data that suggest that the treatment caused a reduction in absences. The bottom panel shows data that suggest that it did not.

Combination Designs

A type of quasi-experimental design that is generally better than either the nonequivalent groups design or the pretest-posttest design is one that combines elements of both. There is a treatment group that is given a pretest, receives a treatment, and then is given a posttest. But at the same time there is a control group that is given a pretest, does not receive the treatment, and then is given a posttest. The question, then, is not simply whether participants who receive the treatment improve but whether they improve more than participants who do not receive the treatment.

Imagine, for example, that students in one school are given a pretest on their attitudes toward drugs, then are exposed to an antidrug program, and finally are given a posttest. Students in a similar school are given the pretest, not exposed to an antidrug program, and finally are given a posttest. Again, if students in the treatment condition become more negative toward drugs, this could be an effect of the treatment, but it could also be a matter of history or maturation. If it really is an effect of the treatment, then students in the treatment condition should become more negative than students in the control condition. But if it is a matter of history (e.g., news of a celebrity drug overdose) or maturation (e.g., improved reasoning), then students in the two conditions would be likely to show similar amounts of change. This type of design does not completely eliminate the possibility of confounding variables, however. Something could occur at one of the schools but not the other (e.g., a student drug overdose), so students at the first school would be affected by it while students at the other school would not.

Finally, if participants in this kind of design are randomly assigned to conditions, it becomes a true experiment rather than a quasi experiment. In fact, it is the kind of experiment that Eysenck called for—and that has now been conducted many times—to demonstrate the effectiveness of psychotherapy.

Key Takeaways

  • Quasi-experimental research involves the manipulation of an independent variable without the random assignment of participants to conditions or orders of conditions. Among the important types are nonequivalent groups designs, pretest-posttest, and interrupted time-series designs.
  • Quasi-experimental research eliminates the directionality problem because it involves the manipulation of the independent variable. It does not eliminate the problem of confounding variables, however, because it does not involve random assignment to conditions. For these reasons, quasi-experimental research is generally higher in internal validity than correlational studies but lower than true experiments.
  • Practice: Imagine that two college professors decide to test the effect of giving daily quizzes on student performance in a statistics course. They decide that Professor A will give quizzes but Professor B will not. They will then compare the performance of students in their two sections on a common final exam. List five other variables that might differ between the two sections that could affect the results.

Discussion: Imagine that a group of obese children is recruited for a study in which their weight is measured, then they participate for 3 months in a program that encourages them to be more active, and finally their weight is measured again. Explain how each of the following might affect the results:

  • regression to the mean
  • spontaneous remission

Cook, T. D., & Campbell, D. T. (1979). Quasi-experimentation: Design & analysis issues in field settings . Boston, MA: Houghton Mifflin.

Eysenck, H. J. (1952). The effects of psychotherapy: An evaluation. Journal of Consulting Psychology, 16 , 319–324.

Posternak, M. A., & Miller, I. (2001). Untreated short-term course of major depression: A meta-analysis of studies using outcomes from studies using wait-list control groups. Journal of Affective Disorders, 66 , 139–146.

Smith, M. L., Glass, G. V., & Miller, T. I. (1980). The benefits of psychotherapy . Baltimore, MD: Johns Hopkins University Press.

Research Methods in Psychology Copyright © 2016 by University of Minnesota is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License , except where otherwise noted.

Experimental and Quasi-Experimental Methods

  • Reference work entry
  • First Online: 01 January 2014
  • Cite this reference work entry

random assignment quasi experimental

  • Roger J. R. Levesque 2  

112 Accesses

Research designs are central to research projects in that they constitute the projects’ basic structure that will permit researchers to address their main research questions. Designs include, for example, the selection of relevant samples or groups, measures, treatments or programs, and methods of assignment. The two key designs that help researchers address whether a program or treatment causes an outcome are the experimental design, which uses random assignment to groups or programs, and quasi-experimental designs, which do not use random assignment (see Shadish et al. 2002 ; Bell 2010 ; Trochim 2006 ). These two methods are important to consider in that even the experimental design may not prove causation, and causation is what researchers often aim to show when they analyze data (e.g., they try to show that an outcome is likely to follow given a certain set of conditions). Still, the general rule tends to be that studies unable to determine causality are classified as...

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

Access this chapter

  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Bell, S. H. (2010). The urban institute research of record: Quasi-experimental methods. Washington, DC: The Urban Institute. Retrieved Nov. 20, 2010, from http://www.urban.org/toolkit/data-methods/quasi-experimental.cfm

Campbell, D. T., & Stanley, J. C. (1966). Experimental and quasi-experimental designs for research . Chicago: Rand McNally.

Google Scholar  

Harris, A. D., McGregor, J. C., Perencevich, E. N., Furuno, J. P., Zhu, J., Peterson, D. E., & Finkelstein, J. (2006). The use and interpretation of quasi-experimental studies in medical informatics. The Journal of American Medical Informatics Association, 13 , 16–23.

Shadish, W. R., Cook, T. D., & Campbell, T. D. (2002). Experimental and quasi-experimental designs for generalized causal inference . Boston: Houghton-Mifflin.

Trochim, W. M. (2006). The research methods knowledge base (2nd ed.). Cincinnati: Atomic Dog. Retrieved Nov. 20, 2011, from http://www.socialresearchmethods.net/kb/

Download references

Author information

Authors and affiliations.

Indiana University, 302 Sycamore Hall, Bloomington, IN, 47405, USA

Roger J. R. Levesque

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Roger J. R. Levesque .

Editor information

Editors and affiliations, rights and permissions.

Reprints and permissions

Copyright information

© 2011 Springer Science+Business Media, LLC

About this entry

Cite this entry.

Levesque, R.J.R. (2011). Experimental and Quasi-Experimental Methods. In: Levesque, R.J.R. (eds) Encyclopedia of Adolescence. Springer, New York, NY. https://doi.org/10.1007/978-1-4419-1695-2_655

Download citation

DOI : https://doi.org/10.1007/978-1-4419-1695-2_655

Published : 12 August 2014

Publisher Name : Springer, New York, NY

Print ISBN : 978-1-4419-1694-5

Online ISBN : 978-1-4419-1695-2

eBook Packages : Behavioral Science

Share this entry

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research

Research Methodologies Guide

  • Action Research
  • Bibliometrics
  • Case Studies
  • Content Analysis
  • Digital Scholarship This link opens in a new window
  • Documentary
  • Ethnography
  • Focus Groups
  • Grounded Theory
  • Life Histories/Autobiographies
  • Longitudinal
  • Participant Observation
  • Qualitative Research (General)

Quasi-Experimental Design

  • Usability Studies

Quasi-Experimental Design is a unique research methodology because it is characterized by what is lacks. For example, Abraham & MacDonald (2011) state:

" Quasi-experimental research is similar to experimental research in that there is manipulation of an independent variable. It differs from experimental research because either there is no control group, no random selection, no random assignment, and/or no active manipulation. "

This type of research is often performed in cases where a control group cannot be created or random selection cannot be performed. This is often the case in certain medical and psychological studies. 

For more information on quasi-experimental design, review the resources below: 

Where to Start

Below are listed a few tools and online guides that can help you start your Quasi-experimental research. These include free online resources and resources available only through ISU Library.

  • Quasi-Experimental Research Designs by Bruce A. Thyer This pocket guide describes the logic, design, and conduct of the range of quasi-experimental designs, encompassing pre-experiments, quasi-experiments making use of a control or comparison group, and time-series designs. An introductory chapter describes the valuable role these types of studies have played in social work, from the 1930s to the present. Subsequent chapters delve into each design type's major features, the kinds of questions it is capable of answering, and its strengths and limitations.
  • Experimental and Quasi-Experimental Designs for Research by Donald T. Campbell; Julian C. Stanley. Call Number: Q175 C152e Written 1967 but still used heavily today, this book examines research designs for experimental and quasi-experimental research, with examples and judgments about each design's validity.

Online Resources

  • Quasi-Experimental Design From the Web Center for Social Research Methods, this is a very good overview of quasi-experimental design.
  • Experimental and Quasi-Experimental Research From Colorado State University.
  • Quasi-experimental design--Wikipedia, the free encyclopedia Wikipedia can be a useful place to start your research- check the citations at the bottom of the article for more information.
  • << Previous: Qualitative Research (General)
  • Next: Sampling >>
  • Last Updated: Dec 19, 2023 2:12 PM
  • URL: https://instr.iastate.libguides.com/researchmethods
  • Privacy Policy

Research Method

Home » Quasi-Experimental Research Design – Types, Methods

Quasi-Experimental Research Design – Types, Methods

Table of Contents

Quasi-Experimental Design

Quasi-Experimental Design

Quasi-experimental design is a research method that seeks to evaluate the causal relationships between variables, but without the full control over the independent variable(s) that is available in a true experimental design.

In a quasi-experimental design, the researcher uses an existing group of participants that is not randomly assigned to the experimental and control groups. Instead, the groups are selected based on pre-existing characteristics or conditions, such as age, gender, or the presence of a certain medical condition.

Types of Quasi-Experimental Design

There are several types of quasi-experimental designs that researchers use to study causal relationships between variables. Here are some of the most common types:

Non-Equivalent Control Group Design

This design involves selecting two groups of participants that are similar in every way except for the independent variable(s) that the researcher is testing. One group receives the treatment or intervention being studied, while the other group does not. The two groups are then compared to see if there are any significant differences in the outcomes.

Interrupted Time-Series Design

This design involves collecting data on the dependent variable(s) over a period of time, both before and after an intervention or event. The researcher can then determine whether there was a significant change in the dependent variable(s) following the intervention or event.

Pretest-Posttest Design

This design involves measuring the dependent variable(s) before and after an intervention or event, but without a control group. This design can be useful for determining whether the intervention or event had an effect, but it does not allow for control over other factors that may have influenced the outcomes.

Regression Discontinuity Design

This design involves selecting participants based on a specific cutoff point on a continuous variable, such as a test score. Participants on either side of the cutoff point are then compared to determine whether the intervention or event had an effect.

Natural Experiments

This design involves studying the effects of an intervention or event that occurs naturally, without the researcher’s intervention. For example, a researcher might study the effects of a new law or policy that affects certain groups of people. This design is useful when true experiments are not feasible or ethical.

Data Analysis Methods

Here are some data analysis methods that are commonly used in quasi-experimental designs:

Descriptive Statistics

This method involves summarizing the data collected during a study using measures such as mean, median, mode, range, and standard deviation. Descriptive statistics can help researchers identify trends or patterns in the data, and can also be useful for identifying outliers or anomalies.

Inferential Statistics

This method involves using statistical tests to determine whether the results of a study are statistically significant. Inferential statistics can help researchers make generalizations about a population based on the sample data collected during the study. Common statistical tests used in quasi-experimental designs include t-tests, ANOVA, and regression analysis.

Propensity Score Matching

This method is used to reduce bias in quasi-experimental designs by matching participants in the intervention group with participants in the control group who have similar characteristics. This can help to reduce the impact of confounding variables that may affect the study’s results.

Difference-in-differences Analysis

This method is used to compare the difference in outcomes between two groups over time. Researchers can use this method to determine whether a particular intervention has had an impact on the target population over time.

Interrupted Time Series Analysis

This method is used to examine the impact of an intervention or treatment over time by comparing data collected before and after the intervention or treatment. This method can help researchers determine whether an intervention had a significant impact on the target population.

Regression Discontinuity Analysis

This method is used to compare the outcomes of participants who fall on either side of a predetermined cutoff point. This method can help researchers determine whether an intervention had a significant impact on the target population.

Steps in Quasi-Experimental Design

Here are the general steps involved in conducting a quasi-experimental design:

  • Identify the research question: Determine the research question and the variables that will be investigated.
  • Choose the design: Choose the appropriate quasi-experimental design to address the research question. Examples include the pretest-posttest design, non-equivalent control group design, regression discontinuity design, and interrupted time series design.
  • Select the participants: Select the participants who will be included in the study. Participants should be selected based on specific criteria relevant to the research question.
  • Measure the variables: Measure the variables that are relevant to the research question. This may involve using surveys, questionnaires, tests, or other measures.
  • Implement the intervention or treatment: Implement the intervention or treatment to the participants in the intervention group. This may involve training, education, counseling, or other interventions.
  • Collect data: Collect data on the dependent variable(s) before and after the intervention. Data collection may also include collecting data on other variables that may impact the dependent variable(s).
  • Analyze the data: Analyze the data collected to determine whether the intervention had a significant impact on the dependent variable(s).
  • Draw conclusions: Draw conclusions about the relationship between the independent and dependent variables. If the results suggest a causal relationship, then appropriate recommendations may be made based on the findings.

Quasi-Experimental Design Examples

Here are some examples of real-time quasi-experimental designs:

  • Evaluating the impact of a new teaching method: In this study, a group of students are taught using a new teaching method, while another group is taught using the traditional method. The test scores of both groups are compared before and after the intervention to determine whether the new teaching method had a significant impact on student performance.
  • Assessing the effectiveness of a public health campaign: In this study, a public health campaign is launched to promote healthy eating habits among a targeted population. The behavior of the population is compared before and after the campaign to determine whether the intervention had a significant impact on the target behavior.
  • Examining the impact of a new medication: In this study, a group of patients is given a new medication, while another group is given a placebo. The outcomes of both groups are compared to determine whether the new medication had a significant impact on the targeted health condition.
  • Evaluating the effectiveness of a job training program : In this study, a group of unemployed individuals is enrolled in a job training program, while another group is not enrolled in any program. The employment rates of both groups are compared before and after the intervention to determine whether the training program had a significant impact on the employment rates of the participants.
  • Assessing the impact of a new policy : In this study, a new policy is implemented in a particular area, while another area does not have the new policy. The outcomes of both areas are compared before and after the intervention to determine whether the new policy had a significant impact on the targeted behavior or outcome.

Applications of Quasi-Experimental Design

Here are some applications of quasi-experimental design:

  • Educational research: Quasi-experimental designs are used to evaluate the effectiveness of educational interventions, such as new teaching methods, technology-based learning, or educational policies.
  • Health research: Quasi-experimental designs are used to evaluate the effectiveness of health interventions, such as new medications, public health campaigns, or health policies.
  • Social science research: Quasi-experimental designs are used to investigate the impact of social interventions, such as job training programs, welfare policies, or criminal justice programs.
  • Business research: Quasi-experimental designs are used to evaluate the impact of business interventions, such as marketing campaigns, new products, or pricing strategies.
  • Environmental research: Quasi-experimental designs are used to evaluate the impact of environmental interventions, such as conservation programs, pollution control policies, or renewable energy initiatives.

When to use Quasi-Experimental Design

Here are some situations where quasi-experimental designs may be appropriate:

  • When the research question involves investigating the effectiveness of an intervention, policy, or program : In situations where it is not feasible or ethical to randomly assign participants to intervention and control groups, quasi-experimental designs can be used to evaluate the impact of the intervention on the targeted outcome.
  • When the sample size is small: In situations where the sample size is small, it may be difficult to randomly assign participants to intervention and control groups. Quasi-experimental designs can be used to investigate the impact of an intervention without requiring a large sample size.
  • When the research question involves investigating a naturally occurring event : In some situations, researchers may be interested in investigating the impact of a naturally occurring event, such as a natural disaster or a major policy change. Quasi-experimental designs can be used to evaluate the impact of the event on the targeted outcome.
  • When the research question involves investigating a long-term intervention: In situations where the intervention or program is long-term, it may be difficult to randomly assign participants to intervention and control groups for the entire duration of the intervention. Quasi-experimental designs can be used to evaluate the impact of the intervention over time.
  • When the research question involves investigating the impact of a variable that cannot be manipulated : In some situations, it may not be possible or ethical to manipulate a variable of interest. Quasi-experimental designs can be used to investigate the relationship between the variable and the targeted outcome.

Purpose of Quasi-Experimental Design

The purpose of quasi-experimental design is to investigate the causal relationship between two or more variables when it is not feasible or ethical to conduct a randomized controlled trial (RCT). Quasi-experimental designs attempt to emulate the randomized control trial by mimicking the control group and the intervention group as much as possible.

The key purpose of quasi-experimental design is to evaluate the impact of an intervention, policy, or program on a targeted outcome while controlling for potential confounding factors that may affect the outcome. Quasi-experimental designs aim to answer questions such as: Did the intervention cause the change in the outcome? Would the outcome have changed without the intervention? And was the intervention effective in achieving its intended goals?

Quasi-experimental designs are useful in situations where randomized controlled trials are not feasible or ethical. They provide researchers with an alternative method to evaluate the effectiveness of interventions, policies, and programs in real-life settings. Quasi-experimental designs can also help inform policy and practice by providing valuable insights into the causal relationships between variables.

Overall, the purpose of quasi-experimental design is to provide a rigorous method for evaluating the impact of interventions, policies, and programs while controlling for potential confounding factors that may affect the outcome.

Advantages of Quasi-Experimental Design

Quasi-experimental designs have several advantages over other research designs, such as:

  • Greater external validity : Quasi-experimental designs are more likely to have greater external validity than laboratory experiments because they are conducted in naturalistic settings. This means that the results are more likely to generalize to real-world situations.
  • Ethical considerations: Quasi-experimental designs often involve naturally occurring events, such as natural disasters or policy changes. This means that researchers do not need to manipulate variables, which can raise ethical concerns.
  • More practical: Quasi-experimental designs are often more practical than experimental designs because they are less expensive and easier to conduct. They can also be used to evaluate programs or policies that have already been implemented, which can save time and resources.
  • No random assignment: Quasi-experimental designs do not require random assignment, which can be difficult or impossible in some cases, such as when studying the effects of a natural disaster. This means that researchers can still make causal inferences, although they must use statistical techniques to control for potential confounding variables.
  • Greater generalizability : Quasi-experimental designs are often more generalizable than experimental designs because they include a wider range of participants and conditions. This can make the results more applicable to different populations and settings.

Limitations of Quasi-Experimental Design

There are several limitations associated with quasi-experimental designs, which include:

  • Lack of Randomization: Quasi-experimental designs do not involve randomization of participants into groups, which means that the groups being studied may differ in important ways that could affect the outcome of the study. This can lead to problems with internal validity and limit the ability to make causal inferences.
  • Selection Bias: Quasi-experimental designs may suffer from selection bias because participants are not randomly assigned to groups. Participants may self-select into groups or be assigned based on pre-existing characteristics, which may introduce bias into the study.
  • History and Maturation: Quasi-experimental designs are susceptible to history and maturation effects, where the passage of time or other events may influence the outcome of the study.
  • Lack of Control: Quasi-experimental designs may lack control over extraneous variables that could influence the outcome of the study. This can limit the ability to draw causal inferences from the study.
  • Limited Generalizability: Quasi-experimental designs may have limited generalizability because the results may only apply to the specific population and context being studied.

About the author

' src=

Muhammad Hassan

Researcher, Academic Writer, Web developer

You may also like

Questionnaire

Questionnaire – Definition, Types, and Examples

Case Study Research

Case Study – Methods, Examples and Guide

Observational Research

Observational Research – Methods and Guide

Quantitative Research

Quantitative Research – Methods, Types and...

Qualitative Research Methods

Qualitative Research Methods

Explanatory Research

Explanatory Research – Types, Methods, Guide

The prefix  quasi  means “resembling.” Thus quasi-experimental research is research that resembles experimental research but is not true experimental research. Recall with a true between-groups experiment, random assignment to conditions is used to ensure the groups are equivalent and with a true within-subjects design counterbalancing is used to guard against order effects. Quasi-experiments are missing one of these safeguards. Although an independent variable is manipulated, either a control group is missing or participants are not randomly assigned to conditions (Cook & Campbell, 1979) [1] .

Because the independent variable is manipulated before the dependent variable is measured, quasi-experimental research eliminates the directionality problem associated with non-experimental research. But because either counterbalancing techniques are not used or participants are not randomly assigned to conditions—making it likely that there are other differences between conditions—quasi-experimental research does not eliminate the problem of confounding variables. In terms of internal validity, therefore, quasi-experiments are generally somewhere between non-experimental studies and true experiments.

Quasi-experiments are most likely to be conducted in field settings in which random assignment is difficult or impossible. They are often conducted to evaluate the effectiveness of a treatment—perhaps a type of psychotherapy or an educational intervention. There are many different kinds of quasi-experiments, but we will discuss just a few of the most common ones in this chapter. 

  • Cook, T. D., & Campbell, D. T. (1979). Quasi-experimentation: Design & analysis issues in field settings . Boston, MA: Houghton Mifflin. ↵

Creative Commons License

Share This Book

  • Increase Font Size

Logo for BCcampus Open Publishing

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

Chapter 7: Nonexperimental Research

Quasi-Experimental Research

Learning Objectives

  • Explain what quasi-experimental research is and distinguish it clearly from both experimental and correlational research.
  • Describe three different types of quasi-experimental research designs (nonequivalent groups, pretest-posttest, and interrupted time series) and identify examples of each one.

The prefix  quasi  means “resembling.” Thus quasi-experimental research is research that resembles experimental research but is not true experimental research. Although the independent variable is manipulated, participants are not randomly assigned to conditions or orders of conditions (Cook & Campbell, 1979). [1] Because the independent variable is manipulated before the dependent variable is measured, quasi-experimental research eliminates the directionality problem. But because participants are not randomly assigned—making it likely that there are other differences between conditions—quasi-experimental research does not eliminate the problem of confounding variables. In terms of internal validity, therefore, quasi-experiments are generally somewhere between correlational studies and true experiments.

Quasi-experiments are most likely to be conducted in field settings in which random assignment is difficult or impossible. They are often conducted to evaluate the effectiveness of a treatment—perhaps a type of psychotherapy or an educational intervention. There are many different kinds of quasi-experiments, but we will discuss just a few of the most common ones here.

Nonequivalent Groups Design

Recall that when participants in a between-subjects experiment are randomly assigned to conditions, the resulting groups are likely to be quite similar. In fact, researchers consider them to be equivalent. When participants are not randomly assigned to conditions, however, the resulting groups are likely to be dissimilar in some ways. For this reason, researchers consider them to be nonequivalent. A  nonequivalent groups design , then, is a between-subjects design in which participants have not been randomly assigned to conditions.

Imagine, for example, a researcher who wants to evaluate a new method of teaching fractions to third graders. One way would be to conduct a study with a treatment group consisting of one class of third-grade students and a control group consisting of another class of third-grade students. This design would be a nonequivalent groups design because the students are not randomly assigned to classes by the researcher, which means there could be important differences between them. For example, the parents of higher achieving or more motivated students might have been more likely to request that their children be assigned to Ms. Williams’s class. Or the principal might have assigned the “troublemakers” to Mr. Jones’s class because he is a stronger disciplinarian. Of course, the teachers’ styles, and even the classroom environments, might be very different and might cause different levels of achievement or motivation among the students. If at the end of the study there was a difference in the two classes’ knowledge of fractions, it might have been caused by the difference between the teaching methods—but it might have been caused by any of these confounding variables.

Of course, researchers using a nonequivalent groups design can take steps to ensure that their groups are as similar as possible. In the present example, the researcher could try to select two classes at the same school, where the students in the two classes have similar scores on a standardized math test and the teachers are the same sex, are close in age, and have similar teaching styles. Taking such steps would increase the internal validity of the study because it would eliminate some of the most important confounding variables. But without true random assignment of the students to conditions, there remains the possibility of other important confounding variables that the researcher was not able to control.

Pretest-Posttest Design

In a  pretest-posttest design , the dependent variable is measured once before the treatment is implemented and once after it is implemented. Imagine, for example, a researcher who is interested in the effectiveness of an antidrug education program on elementary school students’ attitudes toward illegal drugs. The researcher could measure the attitudes of students at a particular elementary school during one week, implement the antidrug program during the next week, and finally, measure their attitudes again the following week. The pretest-posttest design is much like a within-subjects experiment in which each participant is tested first under the control condition and then under the treatment condition. It is unlike a within-subjects experiment, however, in that the order of conditions is not counterbalanced because it typically is not possible for a participant to be tested in the treatment condition first and then in an “untreated” control condition.

If the average posttest score is better than the average pretest score, then it makes sense to conclude that the treatment might be responsible for the improvement. Unfortunately, one often cannot conclude this with a high degree of certainty because there may be other explanations for why the posttest scores are better. One category of alternative explanations goes under the name of  history . Other things might have happened between the pretest and the posttest. Perhaps an antidrug program aired on television and many of the students watched it, or perhaps a celebrity died of a drug overdose and many of the students heard about it. Another category of alternative explanations goes under the name of  maturation . Participants might have changed between the pretest and the posttest in ways that they were going to anyway because they are growing and learning. If it were a yearlong program, participants might become less impulsive or better reasoners and this might be responsible for the change.

Another alternative explanation for a change in the dependent variable in a pretest-posttest design is  regression to the mean . This refers to the statistical fact that an individual who scores extremely on a variable on one occasion will tend to score less extremely on the next occasion. For example, a bowler with a long-term average of 150 who suddenly bowls a 220 will almost certainly score lower in the next game. Her score will “regress” toward her mean score of 150. Regression to the mean can be a problem when participants are selected for further study  because  of their extreme scores. Imagine, for example, that only students who scored especially low on a test of fractions are given a special training program and then retested. Regression to the mean all but guarantees that their scores will be higher even if the training program has no effect. A closely related concept—and an extremely important one in psychological research—is  spontaneous remission . This is the tendency for many medical and psychological problems to improve over time without any form of treatment. The common cold is a good example. If one were to measure symptom severity in 100 common cold sufferers today, give them a bowl of chicken soup every day, and then measure their symptom severity again in a week, they would probably be much improved. This does not mean that the chicken soup was responsible for the improvement, however, because they would have been much improved without any treatment at all. The same is true of many psychological problems. A group of severely depressed people today is likely to be less depressed on average in 6 months. In reviewing the results of several studies of treatments for depression, researchers Michael Posternak and Ivan Miller found that participants in waitlist control conditions improved an average of 10 to 15% before they received any treatment at all (Posternak & Miller, 2001) [2] . Thus one must generally be very cautious about inferring causality from pretest-posttest designs.

Does Psychotherapy Work?

Early studies on the effectiveness of psychotherapy tended to use pretest-posttest designs. In a classic 1952 article, researcher Hans Eysenck summarized the results of 24 such studies showing that about two thirds of patients improved between the pretest and the posttest (Eysenck, 1952) [3] . But Eysenck also compared these results with archival data from state hospital and insurance company records showing that similar patients recovered at about the same rate  without  receiving psychotherapy. This parallel suggested to Eysenck that the improvement that patients showed in the pretest-posttest studies might be no more than spontaneous remission. Note that Eysenck did not conclude that psychotherapy was ineffective. He merely concluded that there was no evidence that it was, and he wrote of “the necessity of properly planned and executed experimental studies into this important field” (p. 323). You can read the entire article here: Classics in the History of Psychology .

Fortunately, many other researchers took up Eysenck’s challenge, and by 1980 hundreds of experiments had been conducted in which participants were randomly assigned to treatment and control conditions, and the results were summarized in a classic book by Mary Lee Smith, Gene Glass, and Thomas Miller (Smith, Glass, & Miller, 1980) [4] . They found that overall psychotherapy was quite effective, with about 80% of treatment participants improving more than the average control participant. Subsequent research has focused more on the conditions under which different types of psychotherapy are more or less effective.

Interrupted Time Series Design

A variant of the pretest-posttest design is the  interrupted time-series design . A time series is a set of measurements taken at intervals over a period of time. For example, a manufacturing company might measure its workers’ productivity each week for a year. In an interrupted time series-design, a time series like this one is “interrupted” by a treatment. In one classic example, the treatment was the reduction of the work shifts in a factory from 10 hours to 8 hours (Cook & Campbell, 1979) [5] . Because productivity increased rather quickly after the shortening of the work shifts, and because it remained elevated for many months afterward, the researcher concluded that the shortening of the shifts caused the increase in productivity. Notice that the interrupted time-series design is like a pretest-posttest design in that it includes measurements of the dependent variable both before and after the treatment. It is unlike the pretest-posttest design, however, in that it includes multiple pretest and posttest measurements.

Figure 7.3 shows data from a hypothetical interrupted time-series study. The dependent variable is the number of student absences per week in a research methods course. The treatment is that the instructor begins publicly taking attendance each day so that students know that the instructor is aware of who is present and who is absent. The top panel of  Figure 7.3 shows how the data might look if this treatment worked. There is a consistently high number of absences before the treatment, and there is an immediate and sustained drop in absences after the treatment. The bottom panel of  Figure 7.3 shows how the data might look if this treatment did not work. On average, the number of absences after the treatment is about the same as the number before. This figure also illustrates an advantage of the interrupted time-series design over a simpler pretest-posttest design. If there had been only one measurement of absences before the treatment at Week 7 and one afterward at Week 8, then it would have looked as though the treatment were responsible for the reduction. The multiple measurements both before and after the treatment suggest that the reduction between Weeks 7 and 8 is nothing more than normal week-to-week variation.

Image description available

Combination Designs

A type of quasi-experimental design that is generally better than either the nonequivalent groups design or the pretest-posttest design is one that combines elements of both. There is a treatment group that is given a pretest, receives a treatment, and then is given a posttest. But at the same time there is a control group that is given a pretest, does  not  receive the treatment, and then is given a posttest. The question, then, is not simply whether participants who receive the treatment improve but whether they improve  more  than participants who do not receive the treatment.

Imagine, for example, that students in one school are given a pretest on their attitudes toward drugs, then are exposed to an antidrug program, and finally are given a posttest. Students in a similar school are given the pretest, not exposed to an antidrug program, and finally are given a posttest. Again, if students in the treatment condition become more negative toward drugs, this change in attitude could be an effect of the treatment, but it could also be a matter of history or maturation. If it really is an effect of the treatment, then students in the treatment condition should become more negative than students in the control condition. But if it is a matter of history (e.g., news of a celebrity drug overdose) or maturation (e.g., improved reasoning), then students in the two conditions would be likely to show similar amounts of change. This type of design does not completely eliminate the possibility of confounding variables, however. Something could occur at one of the schools but not the other (e.g., a student drug overdose), so students at the first school would be affected by it while students at the other school would not.

Finally, if participants in this kind of design are randomly assigned to conditions, it becomes a true experiment rather than a quasi experiment. In fact, it is the kind of experiment that Eysenck called for—and that has now been conducted many times—to demonstrate the effectiveness of psychotherapy.

Key Takeaways

  • Quasi-experimental research involves the manipulation of an independent variable without the random assignment of participants to conditions or orders of conditions. Among the important types are nonequivalent groups designs, pretest-posttest, and interrupted time-series designs.
  • Quasi-experimental research eliminates the directionality problem because it involves the manipulation of the independent variable. It does not eliminate the problem of confounding variables, however, because it does not involve random assignment to conditions. For these reasons, quasi-experimental research is generally higher in internal validity than correlational studies but lower than true experiments.
  • Practice: Imagine that two professors decide to test the effect of giving daily quizzes on student performance in a statistics course. They decide that Professor A will give quizzes but Professor B will not. They will then compare the performance of students in their two sections on a common final exam. List five other variables that might differ between the two sections that could affect the results.
  • regression to the mean
  • spontaneous remission

Image Descriptions

Figure 7.3 image description: Two line graphs charting the number of absences per week over 14 weeks. The first 7 weeks are without treatment and the last 7 weeks are with treatment. In the first line graph, there are between 4 to 8 absences each week. After the treatment, the absences drop to 0 to 3 each week, which suggests the treatment worked. In the second line graph, there is no noticeable change in the number of absences per week after the treatment, which suggests the treatment did not work. [Return to Figure 7.3]

  • Cook, T. D., & Campbell, D. T. (1979). Quasi-experimentation: Design & analysis issues in field settings . Boston, MA: Houghton Mifflin. ↵
  • Posternak, M. A., & Miller, I. (2001). Untreated short-term course of major depression: A meta-analysis of studies using outcomes from studies using wait-list control groups. Journal of Affective Disorders, 66 , 139–146. ↵
  • Eysenck, H. J. (1952). The effects of psychotherapy: An evaluation. Journal of Consulting Psychology, 16 , 319–324. ↵
  • Smith, M. L., Glass, G. V., & Miller, T. I. (1980). The benefits of psychotherapy . Baltimore, MD: Johns Hopkins University Press. ↵

A between-subjects design in which participants have not been randomly assigned to conditions.

The dependent variable is measured once before the treatment is implemented and once after it is implemented.

A category of alternative explanations for differences between scores such as events that happened between the pretest and posttest, unrelated to the study.

An alternative explanation that refers to how the participants might have changed between the pretest and posttest in ways that they were going to anyway because they are growing and learning.

The statistical fact that an individual who scores extremely on a variable on one occasion will tend to score less extremely on the next occasion.

The tendency for many medical and psychological problems to improve over time without any form of treatment.

A set of measurements taken at intervals over a period of time that are interrupted by a treatment.

Research Methods in Psychology - 2nd Canadian Edition Copyright © 2015 by Paul C. Price, Rajiv Jhangiani, & I-Chant A. Chiang is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License , except where otherwise noted.

Share This Book

random assignment quasi experimental

Logo for Mavs Open Press

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

8.2 Quasi-experimental and pre-experimental designs

Learning objectives.

  • Identify and describe the various types of quasi-experimental designs
  • Distinguish true experimental designs from quasi-experimental and pre-experimental designs
  • Identify and describe the various types of quasi-experimental and pre-experimental designs

As we discussed in the previous section, time, funding, and ethics may limit a researcher’s ability to conduct a true experiment. For researchers in the medical sciences and social work, conducting a true experiment could require denying needed treatment to clients, which is a clear ethical violation. Even those whose research may not involve the administration of needed medications or treatments may be limited in their ability to conduct a classic experiment. When true experiments are not possible, researchers often use quasi-experimental designs.

Quasi-experimental designs

Quasi-experimental designs are similar to true experiments, but they lack random assignment to experimental and control groups. Quasi-experimental designs have a comparison group that is similar to a control group except assignment to the comparison group is not determined by random assignment. The most basic of these quasi-experimental designs is the nonequivalent comparison groups design (Rubin & Babbie, 2017).  The nonequivalent comparison group design looks a lot like the classic experimental design, except it does not use random assignment. In many cases, these groups may already exist. For example, a researcher might conduct research at two different agency sites, one of which receives the intervention and the other does not. No one was assigned to treatment or comparison groups. Those groupings existed prior to the study. While this method is more convenient for real-world research, it is less likely that that the groups are comparable than if they had been determined by random assignment. Perhaps the treatment group has a characteristic that is unique–for example, higher income or different diagnoses–that make the treatment more effective.

Quasi-experiments are particularly useful in social welfare policy research. Social welfare policy researchers often look for what are termed natural experiments , or situations in which comparable groups are created by differences that already occur in the real world. Natural experiments are a feature of the social world that allows researchers to use the logic of experimental design to investigate the connection between variables. For example, Stratmann and Wille (2016) were interested in the effects of a state healthcare policy called Certificate of Need on the quality of hospitals. They clearly could not randomly assign states to adopt one set of policies or another. Instead, researchers used hospital referral regions, or the areas from which hospitals draw their patients, that spanned across state lines. Because the hospitals were in the same referral region, researchers could be pretty sure that the client characteristics were pretty similar. In this way, they could classify patients in experimental and comparison groups without dictating state policy or telling people where to live.

random assignment quasi experimental

Matching is another approach in quasi-experimental design for assigning people to experimental and comparison groups. It begins with researchers thinking about what variables are important in their study, particularly demographic variables or attributes that might impact their dependent variable. Individual matching involves pairing participants with similar attributes. Then, the matched pair is split—with one participant going to the experimental group and the other to the comparison group. An ex post facto control group , in contrast, is when a researcher matches individuals after the intervention is administered to some participants. Finally, researchers may engage in aggregate matching , in which the comparison group is determined to be similar on important variables.

Time series design

There are many different quasi-experimental designs in addition to the nonequivalent comparison group design described earlier. Describing all of them is beyond the scope of this textbook, but one more design is worth mentioning. The time series design uses multiple observations before and after an intervention. In some cases, experimental and comparison groups are used. In other cases where that is not feasible, a single experimental group is used. By using multiple observations before and after the intervention, the researcher can better understand the true value of the dependent variable in each participant before the intervention starts. Additionally, multiple observations afterwards allow the researcher to see whether the intervention had lasting effects on participants. Time series designs are similar to single-subjects designs, which we will discuss in Chapter 15.

Pre-experimental design

When true experiments and quasi-experiments are not possible, researchers may turn to a pre-experimental design (Campbell & Stanley, 1963).  Pre-experimental designs are called such because they often happen as a pre-cursor to conducting a true experiment.  Researchers want to see if their interventions will have some effect on a small group of people before they seek funding and dedicate time to conduct a true experiment. Pre-experimental designs, thus, are usually conducted as a first step towards establishing the evidence for or against an intervention. However, this type of design comes with some unique disadvantages, which we’ll describe below.

A commonly used type of pre-experiment is the one-group pretest post-test design . In this design, pre- and posttests are both administered, but there is no comparison group to which to compare the experimental group. Researchers may be able to make the claim that participants receiving the treatment experienced a change in the dependent variable, but they cannot begin to claim that the change was the result of the treatment without a comparison group.   Imagine if the students in your research class completed a questionnaire about their level of stress at the beginning of the semester.  Then your professor taught you mindfulness techniques throughout the semester.  At the end of the semester, she administers the stress survey again.  What if levels of stress went up?  Could she conclude that the mindfulness techniques caused stress?  Not without a comparison group!  If there was a comparison group, she would be able to recognize that all students experienced higher stress at the end of the semester than the beginning of the semester, not just the students in her research class.

In cases where the administration of a pretest is cost prohibitive or otherwise not possible, a one- shot case study design might be used. In this instance, no pretest is administered, nor is a comparison group present. If we wished to measure the impact of a natural disaster, such as Hurricane Katrina for example, we might conduct a pre-experiment by identifying  a community that was hit by the hurricane and then measuring the levels of stress in the community.  Researchers using this design must be extremely cautious about making claims regarding the effect of the treatment or stimulus. They have no idea what the levels of stress in the community were before the hurricane hit nor can they compare the stress levels to a community that was not affected by the hurricane.  Nonetheless, this design can be useful for exploratory studies aimed at testing a measures or the feasibility of further study.

In our example of the study of the impact of Hurricane Katrina, a researcher might choose to examine the effects of the hurricane by identifying a group from a community that experienced the hurricane and a comparison group from a similar community that had not been hit by the hurricane. This study design, called a static group comparison , has the advantage of including a comparison group that did not experience the stimulus (in this case, the hurricane). Unfortunately, the design only uses for post-tests, so it is not possible to know if the groups were comparable before the stimulus or intervention.  As you might have guessed from our example, static group comparisons are useful in cases where a researcher cannot control or predict whether, when, or how the stimulus is administered, as in the case of natural disasters.

As implied by the preceding examples where we considered studying the impact of Hurricane Katrina, experiments, quasi-experiments, and pre-experiments do not necessarily need to take place in the controlled setting of a lab. In fact, many applied researchers rely on experiments to assess the impact and effectiveness of various programs and policies. You might recall our discussion of arresting perpetrators of domestic violence in Chapter 2, which is an excellent example of an applied experiment. Researchers did not subject participants to conditions in a lab setting; instead, they applied their stimulus (in this case, arrest) to some subjects in the field and they also had a control group in the field that did not receive the stimulus (and therefore were not arrested).

Key Takeaways

  • Quasi-experimental designs do not use random assignment.
  • Comparison groups are used in quasi-experiments.
  • Matching is a way of improving the comparability of experimental and comparison groups.
  • Quasi-experimental designs and pre-experimental designs are often used when experimental designs are impractical.
  • Quasi-experimental and pre-experimental designs may be easier to carry out, but they lack the rigor of true experiments.
  • Aggregate matching – when the comparison group is determined to be similar to the experimental group along important variables
  • Comparison group – a group in quasi-experimental design that does not receive the experimental treatment; it is similar to a control group except assignment to the comparison group is not determined by random assignment
  • Ex post facto control group – a control group created when a researcher matches individuals after the intervention is administered
  • Individual matching – pairing participants with similar attributes for the purpose of assignment to groups
  • Natural experiments – situations in which comparable groups are created by differences that already occur in the real world
  • Nonequivalent comparison group design – a quasi-experimental design similar to a classic experimental design but without random assignment
  • One-group pretest post-test design – a pre-experimental design that applies an intervention to one group but also includes a pretest
  • One-shot case study – a pre-experimental design that applies an intervention to only one group without a pretest
  • Pre-experimental designs – a variation of experimental design that lacks the rigor of experiments and is often used before a true experiment is conducted
  • Quasi-experimental design – designs lack random assignment to experimental and control groups
  • Static group design – uses an experimental group and a comparison group, without random assignment and pretesting
  • Time series design – a quasi-experimental design that uses multiple observations before and after an intervention

Image attributions

cat and kitten   matching avocado costumes on the couch looking at the camera by Your Best Digs CC-BY-2.0

Foundations of Social Work Research Copyright © 2020 by Rebecca L. Mauldin is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License , except where otherwise noted.

Share This Book

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons
  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Social Sci LibreTexts

12.2: Pre-experimental and quasi-experimental design

  • Last updated
  • Save as PDF
  • Page ID 25667

  • Matthew DeCarlo
  • Radford University via Open Social Work Education

Learning Objectives

  • Identify and describe the various types of quasi-experimental designs
  • Distinguish true experimental designs from quasi-experimental and pre-experimental designs
  • Identify and describe the various types of quasi-experimental and pre-experimental designs

As we discussed in the previous section, time, funding, and ethics may limit a researcher’s ability to conduct a true experiment. For researchers in the medical sciences and social work, conducting a true experiment could require denying needed treatment to clients, which is a clear ethical violation. Even those whose research may not involve the administration of needed medications or treatments may be limited in their ability to conduct a classic experiment. When true experiments are not possible, researchers often use quasi-experimental designs.

Quasi-experimental designs are similar to true experiments, but they lack random assignment to experimental and control groups. The most basic of these quasi-experimental designs is the nonequivalent comparison groups design (Rubin & Babbie, 2017). [1] The nonequivalent comparison group design looks a lot like the classic experimental design, except it does not use random assignment. In many cases, these groups may already exist. For example, a researcher might conduct research at two different agency sites, one of which receives the intervention and the other does not. No one was assigned to treatment or comparison groups. Those groupings existed prior to the study. While this method is more convenient for real-world research, researchers cannot be sure that the groups are comparable. Perhaps the treatment group has a characteristic that is unique–for example, higher income or different diagnoses–that make the treatment more effective.

Quasi-experiments are particularly useful in social welfare policy research. Social welfare policy researchers like me often look for what are termed natural experiments , or situations in which comparable groups are created by differences that already occur in the real world. For example, Stratmann and Wille (2016) [2] were interested in the effects of a state healthcare policy called Certificate of Need on the quality of hospitals. They clearly cannot assign states to adopt one set of policies or another. Instead, researchers used hospital referral regions, or the areas from which hospitals draw their patients, that spanned across state lines. Because the hospitals were in the same referral region, researchers could be pretty sure that the client characteristics were pretty similar. In this way, they could classify patients in experimental and comparison groups without affecting policy or telling people where to live.

There are important examples of policy experiments that use random assignment, including the Oregon Medicaid experiment. In the Oregon Medicaid experiment, the wait list for Oregon was so long, state officials conducted a lottery to see who from the wait list would receive Medicaid (Baicker et al., 2013). [3] Researchers used the lottery as a natural experiment that included random assignment. People selected to be a part of Medicaid were the experimental group and those on the wait list were in the control group. There are some practical complications with using people on a wait list as a control group—most obviously, what happens when people on the wait list are accepted into the program while you’re still collecting data? Natural experiments aren’t a specific kind of experiment like quasi- or pre-experimental designs. Instead, they are more like a feature of the social world that allows researchers to use the logic of experimental design to investigate the connection between variables.

96-1024x682.jpg

Matching is another approach in quasi-experimental design to assigning experimental and comparison groups. Researchers should think about what variables are important in their study, particularly demographic variables or attributes that might impact their dependent variable. Individual matching involves pairing participants with similar attributes. When this is done at the beginning of an experiment, the matched pair is split—with one participant going to the experimental group and the other to the control group. An ex post facto control group , in contrast, is when a researcher matches individuals after the intervention is administered to some participants. Finally, researchers may engage in aggregate matching , in which the comparison group is determined to be similar on important variables.

There are many different quasi-experimental designs in addition to the nonequivalent comparison group design described earlier. Describing all of them is beyond the scope of this textbook, but one more design is worth mentioning. The time series design uses multiple observations before and after an intervention. In some cases, experimental and comparison groups are used. In other cases where that is not feasible, a single experimental group is used. By using multiple observations before and after the intervention, the researcher can better understand the true value of the dependent variable in each participant before the intervention starts. Additionally, multiple observations afterwards allow the researcher to see whether the intervention had lasting effects on participants. Time series designs are similar to single-subjects designs, which we will discuss in Chapter 15.

When true experiments and quasi-experiments are not possible, researchers may turn to a pre-experimental design (Campbell & Stanley, 1963). [4] Pre-experimental designs are called such because they often happen before a true experiment is conducted. Researchers want to see if their interventions will have some effect on a small group of people before they seek funding and dedicate time to conduct a true experiment. Pre-experimental designs, thus, are usually conducted as a first step towards establishing the evidence for or against an intervention. However, this type of design comes with some unique disadvantages, which we’ll describe as we review the pre-experimental designs available.

If we wished to measure the impact of a natural disaster, such as Hurricane Katrina for example, we might conduct a pre-experiment by identifying an experimental group from a community that experienced the hurricane and a control group from a similar community that had not been hit by the hurricane. This study design, called a static group comparison , has the advantage of including a comparison group that did not experience the stimulus (in this case, the hurricane). Unfortunately, it is difficult to know those groups are truly comparable because the experimental and control groups were determined by factors other than random assignment. Additionally, the design would only allow for posttests, unless one were lucky enough to be gathering the data already before Katrina. As you might have guessed from our example, static group comparisons are useful in cases where a researcher cannot control or predict whether, when, or how the stimulus is administered, as in the case of natural disasters.

In cases where the administration of the stimulus is quite costly or otherwise not possible, a one- shot case study design might be used. In this instance, no pretest is administered, nor is a control group present. In our example of the study of the impact of Hurricane Katrina, a researcher using this design would test the impact of Katrina only among a community that was hit by the hurricane and would not seek a comparison group from a community that did not experience the hurricane. Researchers using this design must be extremely cautious about making claims regarding the effect of the stimulus, though the design could be useful for exploratory studies aimed at testing one’s measures or the feasibility of further study.

Finally, if a researcher is unlikely to be able to identify a sample large enough to split into control and experimental groups, or if she simply doesn’t have access to a control group, the researcher might use a one-group pre-/posttest design. In this instance, pre- and posttests are both taken, but there is no control group to which to compare the experimental group. We might be able to study of the impact of Hurricane Katrina using this design if we’d been collecting data on the impacted communities prior to the hurricane. We could then collect similar data after the hurricane. Applying this design involves a bit of serendipity and chance. Without having collected data from impacted communities prior to the hurricane, we would be unable to employ a one- group pre-/posttest design to study Hurricane Katrina’s impact.

As implied by the preceding examples where we considered studying the impact of Hurricane Katrina, experiments do not necessarily need to take place in the controlled setting of a lab. In fact, many applied researchers rely on experiments to assess the impact and effectiveness of various programs and policies. You might recall our discussion of arresting perpetrators of domestic violence in Chapter 6, which is an excellent example of an applied experiment. Researchers did not subject participants to conditions in a lab setting; instead, they applied their stimulus (in this case, arrest) to some subjects in the field and they also had a control group in the field that did not receive the stimulus (and therefore were not arrested).

Key Takeaways

  • Quasi-experimental designs do not use random assignment.
  • Comparison groups are often used in quasi-experiments.
  • Matching is a way of improving the comparability of experimental and comparison groups.
  • Quasi-experimental designs and pre-experimental designs are often used when experimental designs are impractical.
  • Quasi-experimental and pre-experimental designs may be easier to carry out, but they lack the rigor of true experiments.
  • Aggregate matching- when the comparison group is determined to be similar to the experimental group along important variables
  • Ex post facto control group- a control group created when a researcher matches individuals after the intervention is administered
  • Individual matching- pairing participants with similar attributes for the purpose of assignment to groups
  • Natural experiments- situations in which comparable groups are created by differences that already occur in the real world
  • Nonequivalent comparison group design- a quasi-experimental design similar to a classic experimental design but without random assignment
  • One-group pre-/posttest design- a pre-experimental design that applies an intervention to one group but also includes a pretest
  • One-shot case study- a pre-experimental design that applies an intervention to only one group without a pretest
  • Pre-experimental designs- a variation of experimental design that lacks the rigor of experiments and is often used before a true experiment is conducted
  • Quasi-experimental design- designs lack random assignment to experimental and control groups
  • Static group design- uses an experimental group and a comparison group, without random assignment and pretesting
  • Time series design- a quasi-experimental design that uses multiple observations before and after an intervention

Image attributions

cat and kitten matching avocado costumes on the couch looking at the camera by Your Best Digs CC-BY-2.0

  • Rubin, C. & Babbie, S. (2017). Research methods for social work (9th edition) . Boston, MA: Cengage. ↵
  • Stratmann, T. & Wille, D. (2016). Certificate-of-need laws and hospital quality . Mercatus Center at George Mason University, Arlington, VA. Retrieved from: https://www.mercatus.org/system/files/mercatus-stratmann-wille-con-hospital-quality-v1.pdf ↵
  • Baicker, K., Taubman, S. L., Allen, H. L., Bernstein, M., Gruber, J. H., Newhouse, J. P., ... & Finkelstein, A. N. (2013). The Oregon experiment—effects of Medicaid on clinical outcomes. New England Journal of Medicine , 368 (18), 1713-1722. ↵
  • Campbell, D., & Stanley, J. (1963). Experimental and quasi-experimental designs for research . Chicago, IL: Rand McNally. ↵

Experimental vs Quasi-Experimental Design: Which to Choose?

Here’s a table that summarizes the similarities and differences between an experimental and a quasi-experimental study design:

What is a quasi-experimental design?

A quasi-experimental design is a non-randomized study design used to evaluate the effect of an intervention. The intervention can be a training program, a policy change or a medical treatment.

Unlike a true experiment, in a quasi-experimental study the choice of who gets the intervention and who doesn’t is not randomized. Instead, the intervention can be assigned to participants according to their choosing or that of the researcher, or by using any method other than randomness.

Having a control group is not required, but if present, it provides a higher level of evidence for the relationship between the intervention and the outcome.

(for more information, I recommend my other article: Understand Quasi-Experimental Design Through an Example ) .

Examples of quasi-experimental designs include:

  • One-Group Posttest Only Design
  • Static-Group Comparison Design
  • One-Group Pretest-Posttest Design
  • Separate-Sample Pretest-Posttest Design

What is an experimental design?

An experimental design is a randomized study design used to evaluate the effect of an intervention. In its simplest form, the participants will be randomly divided into 2 groups:

  • A treatment group: where participants receive the new intervention which effect we want to study.
  • A control or comparison group: where participants do not receive any intervention at all (or receive some standard intervention).

Randomization ensures that each participant has the same chance of receiving the intervention. Its objective is to equalize the 2 groups, and therefore, any observed difference in the study outcome afterwards will only be attributed to the intervention – i.e. it removes confounding.

(for more information, I recommend my other article: Purpose and Limitations of Random Assignment ).

Examples of experimental designs include:

  • Posttest-Only Control Group Design
  • Pretest-Posttest Control Group Design
  • Solomon Four-Group Design
  • Matched Pairs Design
  • Randomized Block Design

When to choose an experimental design over a quasi-experimental design?

Although many statistical techniques can be used to deal with confounding in a quasi-experimental study, in practice, randomization is still the best tool we have to study causal relationships.

Another problem with quasi-experiments is the natural progression of the disease or the condition under study — When studying the effect of an intervention over time, one should consider natural changes because these can be mistaken with changes in outcome that are caused by the intervention. Having a well-chosen control group helps dealing with this issue.

So, if losing the element of randomness seems like an unwise step down in the hierarchy of evidence, why would we ever want to do it?

This is what we’re going to discuss next.

When to choose a quasi-experimental design over a true experiment?

The issue with randomness is that it cannot be always achievable.

So here are some cases where using a quasi-experimental design makes more sense than using an experimental one:

  • If being in one group is believed to be harmful for the participants , either because the intervention is harmful (ex. randomizing people to smoking), or the intervention has a questionable efficacy, or on the contrary it is believed to be so beneficial that it would be malevolent to put people in the control group (ex. randomizing people to receiving an operation).
  • In cases where interventions act on a group of people in a given location , it becomes difficult to adequately randomize subjects (ex. an intervention that reduces pollution in a given area).
  • When working with small sample sizes , as randomized controlled trials require a large sample size to account for heterogeneity among subjects (i.e. to evenly distribute confounding variables between the intervention and control groups).

Further reading

  • Statistical Software Popularity in 40,582 Research Papers
  • Checking the Popularity of 125 Statistical Tests and Models
  • Objectives of Epidemiology (With Examples)
  • 12 Famous Epidemiologists and Why

UCCS Community

  • Current Students
  • Faculty Staff
  • Alumni & Friends
  • Parents & Families

Schools and Colleges

  • College of Business
  • College of Education
  • College of Engineering and Applied Science
  • College of Letters, Arts & Sciences
  • College of Public Service
  • Graduate School
  • Helen and Arthur E. Johnson Beth-El College of Nursing and Health Sciences

Quick Links

  • Search for Programs & Careers
  • Academic Advising
  • Ent Center for the Arts
  • Kraemer Family Library
  • Military and Veteran Affairs
  • myUCCS Portal
  • Campus Email
  • Microsoft 365
  • Mountain Lion Connect
  • Support Network: Students
  • Support Network: Faculty
  • Account Help

Effect Size Calculators

Dr. Lee A. Becker

  • Content, Part 1
  • Content, Part 2
  • Research Tools

Statistical Analysis of Quasi-Experimental Designs:

I. apriori selection techniques.

Content, part II

I. Overview

Random assignment is used in experimental designs to help assure that different treatment groups are equivalent prior to treatment. With small n 's randomization is messy, the groups may not be equivalent on some important characteristic.

In general, matching is used when you want to make sure that members of the various groups are equivalent on one or more characteristics. If you are want to make absolutely sure that the treatment groups are equivalent on some attribute you can use matched random assignment.

When you can't randomly assign to conditions you can still use matching techniques to try to equate groups on important characteristics. This set of notes makes the distinction between normative group matching and normative group equivalence. In normative group matching you select an exact match from normative comparison group for each participant in the treatment group. In normative group equivalence you select a comparison group that has approximately equivalent characteristics to the treatment group.

II. Matching in Experimental Designs: Matched Random Assignment

In an experimental design, matched random sampling can be used to equate the groups on one or more characteristics. Whitley (in chapter 8) uses an example of matching on IQ.

The Matching Process

Note: Tx = Treatment group, Ctl = Control Group.

Analysis of a Matched Random Assignment Design

If the matching variable is related to the dependent variable, (e.g., IQ is related to almost all studies of memory and learning), then you can incorporate the matching variable as a blocking variable in your analysis of variance. That is, in the 2 x 3 example, the first 6 participants can be entered as IQ block #1, the second 6 participants as IQ block #2. This removes the variance due to IQ from the error term, increasing the power of the study.

The analysis is treated as a repeated measures design where the measures for each block of participants are considered to be repeated measures. For example, in setting up the data for a two-group design (experimental vs. control) the data would look like this:

The analysis would be run as a repeated measures design with group (control vs. experimental) as a within-subjects factor.

If you were interested in analyzing the equivalence of the groups on the IQ score variable you could enter the IQ scores as separate variables.  An analysis of variance of  the IQ scores with treatment group (Treatment vs. Control) as a within-subjects factor should show no mean differences between the two groups. Entering the IQ data would allow you to find the correlation between IQ and performance scores within each treatment group.

One of the problems with this type of analysis is that if any score is missing then the entire block is set to missing.  None of the performance data from Block 4 in Table 2 would be included in the analysis because the performance score is missing for the person in the control group. If you had a 6 cells in your design you would loose the data on all 6 people in a block that had only one missing data point.

I understand that Dr. Klebe has been writing a new data analysis program to take care of this kind of missing data problem.

SPSS Note 

The SPSS syntax commands for running the data in Table 2 as a repeated measures analysis of variance are shown in Table 3.  The SPSS syntax commands for running the data in Table 2 as a paired t test are shown in Table 4. 

III. Matching in Quasi-Experimental Designs: Normative Group Matching

Suppose that you have a quasi-experiment where you want to compare an experimental group (e.g., people who have suffered mild head injury) with a sample from a normative population. Suppose that there are several hundred people in the normative population.

One strategy is to randomly select the same number of people from the normative population as you have in your experimental group. If the demographic characteristics of the normative group approximate those of your experimental group, then this process may be appropriate. But, what if the normative group contains equal numbers of males and females ranging in age from 6 to 102, and people in your experimental condition are all males ranging in age from 18 to 35? Then it is unlikely that the demographic characteristics of the people sampled from the normative group will match those of your experimental group. For that reason, simple random selection is rarely appropriate when sampling from a normative population.

The Normative Group Matching Procedure

Determine the relevant characteristics (e.g., age, gender, SES, etc.) of each person in your experimental group. E.g., Exp person #1 is a 27 year-old male. Then randomly select one of the 27 year-old males from the normative population as a match for Exp person #1. Exp person #2 is a 35 year-old male, then randomly select one of the 35 year-old males as a match for Exp person #2. If you have done randomize normative group matching then the matching variable should be used as a blocking factor in the ANOVA.

If you have a limited number of people in the normative group then you can do caliper matching . In caliper matching you select the matching person based a range of scores, for example, you can caliper match within a range of 3 years. Exp person #1 would be randomly selected from males whose age ranged from 26 to 27 years. If you used a five year caliper for age then for exp person #1 you randomly select a males from those whose age ranged from 25 to 29 years old. You would want a narrower age caliper for children and adolescents than for adults.

This procedure becomes very difficult to accomplish when you try to start matching on more than one variable. Think of the problems of finding exact matches when several variables are used, e.g., an exact match for a 27-year old, white female with an IQ score of 103 and 5 children.

Analysis of a Normative Group Matching Design

The analysis is the same as for a matched random assignment design. If the matching variable is related to the dependent variable, then you can incorporate the matching variable as a blocking variable in your analysis of variance.

III. Matching in Quasi-Experimental Designs: Normative Group Equivalence

Because of the problems in selecting people in a normative group matching design and the potential problems with the data analysis of that design, you may want to make the normative comparison group equivalent on selected demographic characteristics. You might want the same proportion of males and females, and the mean age (and SD) of the normative group should be the same as those in the experimental group. If the ages of the people in the experimental group ranged from 18 to 35, then your normative group might contain an equal number of participants randomly selected from those in the age range from 18 to 35 in the normative population.

Analysis of a Normative Group Equivalence Design

In the case of normative group equivalence there is no special ANOVA procedure as there is in Normative Group Matching. In general, demographic characteristics themselves rarely predict the d.v., so you haven’t lost anything by using the group equivalence method.

A Semantic Caution

The term "matching" implies a one-to-one matching and it implies that you have incorporated that matched variable into your ANOVA design. Please don’t use the term "matching" when you mean mere "equivalence."

  • Link to facebook
  • Link to linkedin
  • Link to twitter
  • Link to youtube
  • Writing Tips

An Introduction to Quasi-Experimental Design

An Introduction to Quasi-Experimental Design

3-minute read

  • 9th January 2022

If you’re a researcher or student in the sciences, you’ll probably come across the term “quasi-experimental design” at some point. But what exactly does it mean?

In this post, we’ll guide you through the different forms of quasi-experimental design and how it compares to true experiments.

What is Quasi-Experimental Design?

Quasi-experimental design (QED) is a research design method that’s useful when regular experimental conditions are impractical or unethical.

Both quasi-experimental designs and true experiments show a cause-and-effect relationship between a dependent and independent variable . Participants in a true experiment are randomly assigned to different treatment groups. The quasi-experimental design, on the other hand, assigns groups based on criteria instead of randomly.

Quasi-Experimental Design Vs. True Experimental Design

The main difference between a quasi-experimental and true experimental design is that in the former, groups aren’t randomly assigned. There are also some other key differences between these research methods.

True experimental design involves:

●     Having control as a researcher over the design of the treatment or program that participants receive (i.e., the independent variable)

●     Control variables as a necessary component

In contrast, a quasi-experimental design involves:

●     The researcher studying groups after they’ve received a treatment or program

●     Control variables as a common element but they aren’t necessary for the experiment to work

Examples of Experimental Design

Perhaps the easiest way to understand quasi-experimental design is to look at how it might be used in practice.

Let’s say you hypothesize that having access to free art lessons will improve the mental health of children from low-income families.

In a true experiment, you’d randomly assign participants to two groups: one that receives free art lessons and another that doesn’t.

However, it’s ethically questionable to deny one group of children access to something that might benefit them.

Find this useful?

Subscribe to our newsletter and get writing tips from our editors straight to your inbox.

Instead, you might decide to compare the data from a community that’s already offered free art classes to these children with that of a community that’s not yet done so.

This second example would be a quasi-experimental design.

Advantages and Disadvantages of Quasi-Experimental Design

Quasi-experimental design has some advantages and disadvantages you’ll need to consider when designing your research.

On the plus side, quasi-experimental design:

●     Has a higher external validity than true experimental design, as it usually involves real-world scenarios

●     Allows you to control for unexpected, confounding variables, resulting in a higher internal validity than other non-experimental methods of research

●     Enables the study of cause-and-effect relationships without the ethical issue of denying a treatment to those who may benefit from it

●     Does not require access to large-scale funding and other practical concerns, as the treatment has already been issued by others

The disadvantages of quasi-experimental design, however, include:

●     Lower internal validity than found in true experiments, as it’s more difficult to account for all confounding variables without using random assignment

●     The necessary data required for research potentially being inaccurate, outdated, or difficult to access

Expert Proofreading for Researchers

We hope our guide has helped you understand the basics of quasi-experimental design.

If you need help with your research paper , our expert proofreaders are available 24/7. Try us out by submitting a free sample document today.

Share this article:

Post A New Comment

Got content that needs a quick turnaround? Let us polish your work. Explore our editorial business services.

How to insert a text box in a google doc.

Google Docs is a powerful collaborative tool, and mastering its features can significantly enhance your...

2-minute read

How to Cite the CDC in APA

If you’re writing about health issues, you might need to reference the Centers for Disease...

5-minute read

Six Product Description Generator Tools for Your Product Copy

Introduction If you’re involved with ecommerce, you’re likely familiar with the often painstaking process of...

What Is a Content Editor?

Are you interested in learning more about the role of a content editor and the...

4-minute read

The Benefits of Using an Online Proofreading Service

Proofreading is important to ensure your writing is clear and concise for your readers. Whether...

6 Online AI Presentation Maker Tools

Creating presentations can be time-consuming and frustrating. Trying to construct a visually appealing and informative...

Logo Harvard University

Make sure your writing is the best it can be with our expert English proofreading and editing.

U.S. flag

An official website of the United States government

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

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

  • Publications
  • Account settings

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

  • Advanced Search
  • Journal List
  • HHS Author Manuscripts

Logo of nihpa

Experimental and Quasi-Experimental Designs in Implementation Research

Christopher j. miller.

a VA Boston Healthcare System, Center for Healthcare Organization and Implementation Research (CHOIR), United States Department of Veterans Affairs, Boston, MA, USA

b Department of Psychiatry, Harvard Medical School, Boston, MA, USA

Shawna N. Smith

c Department of Psychiatry, University of Michigan Medical School, Ann Arbor, MI, USA

d Survey Research Center, Institute for Social Research, University of Michigan, Ann Arbor, MI, USA

Marianne Pugatch

Implementation science is focused on maximizing the adoption, appropriate use, and sustainability of effective clinical practices in real world clinical settings. Many implementation science questions can be feasibly answered by fully experimental designs, typically in the form of randomized controlled trials (RCTs). Implementation-focused RCTs, however, usually differ from traditional efficacy- or effectiveness-oriented RCTs on key parameters. Other implementation science questions are more suited to quasi-experimental designs, which are intended to estimate the effect of an intervention in the absence of randomization. These designs include pre-post designs with a non-equivalent control group, interrupted time series (ITS), and stepped wedges, the last of which require all participants to receive the intervention, but in a staggered fashion. In this article we review the use of experimental designs in implementation science, including recent methodological advances for implementation studies. We also review the use of quasi-experimental designs in implementation science, and discuss the strengths and weaknesses of these approaches. This article is therefore meant to be a practical guide for researchers who are interested in selecting the most appropriate study design to answer relevant implementation science questions, and thereby increase the rate at which effective clinical practices are adopted, spread, and sustained.

1. Background

The first documented clinical trial was conducted in 1747 by James Lind, a royal navy physician, who tested the hypothesis that citrus fruit could cure scurvy. Since then, based on foundational work by Fisher and others (1935), the randomized controlled trial (RCT) has emerged as the gold standard for testing the efficacy of treatment versus a control condition for individual patients. Randomization of patients is seen as a crucial to reducing the impact of measured or unmeasured confounding variables, in turn allowing researchers to draw conclusions regarding causality in clinical trials.

As described elsewhere in this special issue, implementation science is ultimately focused on maximizing the adoption, appropriate use, and sustainability of effective clinical practices in real world clinical settings. As such, some implementation science questions may be addressed by experimental designs. For our purposes here, we use the term “experimental” to refer to designs that feature two essential ingredients: first, manipulation of an independent variable; and second, random assignment of subjects. This corresponds to the definition of randomized experiments originally championed by Fisher (1925) . From this perspective, experimental designs usually take the form of RCTs—but implementation- oriented RCTs typically differ in important ways from traditional efficacy- or effectiveness-oriented RCTs. Other implementation science questions require different methodologies entirely: specifically, several forms of quasi-experimental designs may be used for implementation research in situations where an RCT would be inappropriate. These designs are intended to estimate the effect of an intervention despite a lack of randomization. Quasi-experimental designs include pre-post designs with a nonequivalent control group, interrupted time series (ITS), and stepped wedge designs. Stepped wedges are studies in which all participants receive the intervention, but in a staggered fashion. It is important to note that quasi-experimental designs are not unique to implementation science. As we will discuss below, however, each of them has strengths that make them particularly useful in certain implementation science contexts.

Our goal for this manuscript is two-fold. First, we will summarize the use of experimental designs in implementation science. This will include discussion of ways that implementation-focused RCTs may differ from efficacy- or effectiveness-oriented RCTs. Second, we will summarize the use of quasi-experimental designs in implementation research. This will include discussion of the strengths and weaknesses of these types of approaches in answering implementation research questions. For both experimental and quasi-experimental designs, we will discuss a recent implementation study as an illustrative example of one approach.

1. Experimental Designs in Implementation Science

RCTs in implementation science share the same basic structure as efficacy- or effectiveness-oriented RCTs, but typically feature important distinctions. In this section we will start by reviewing key factors that separate implementation RCTs from more traditional efficacy- or effectiveness-oriented RCTs. We will then discuss optimization trials, which are a type of experimental design that is especially useful for certain implementation science questions. We will then briefly turn our attention to single subject experimental designs (SSEDs) and on-off-on (ABA) designs.

The first common difference that sets apart implementation RCTs from more traditional clinical trials is the primary research question they aim to address. For most implementation trials, the primary research question is not the extent to which a particular treatment or evidence-based practice is more effective than a comparison condition, but instead the extent to which a given implementation strategy is more effective than a comparison condition. For more detail on this pivotal issue, see Drs. Bauer and Kirchner in this special issue.

Second, as a corollary of this point, implementation RCTs typically feature different outcome measures than efficacy or effectiveness RCTs, with an emphasis on the extent to which a health intervention was successfully implemented rather than an evaluation of the health effects of that intervention ( Proctor et al., 2011 ). For example, typical implementation outcomes might include the number of patients who receive the intervention, or the number of providers who administer the intervention as intended. A variety of evaluation-oriented implementation frameworks may guide the choices of such measures (e.g. RE-AIM; Gaglio et al., 2013 ; Glasgow et al., 1999 ). Hybrid implementation-effectiveness studies attend to both effectiveness and implementation outcomes ( Curran et al., 2012 ); these designs are also covered in more detail elsewhere in this issue (Landes, this issue).

Third, given their focus, implementation RCTs are frequently cluster-randomized (i.e. with sites or clinics as the unit of randomization, and patients nested within those sites or clinics). For example, consider a hypothetical RCT that aims to evaluate the implementation of a training program for cognitive behavioral therapy (CBT) in community clinics. Randomizing at the patient level for such a trial would be inappropriate due to the risk of contamination, as providers trained in CBT might reasonably be expected to incorporate CBT principles into their treatment even to patients assigned to the control condition. Randomizing at the provider level would also risk contamination, as providers trained in CBT might discuss this treatment approach with their colleagues. Thus, many implementation trials are cluster randomized at the site or clinic level. While such clustering minimizes the risk of contamination, it can unfortunately create commensurate problems with confounding, especially for trials with very few sites to randomize. Stratification may be used to at least partially address confounding issues in cluster- randomized and more traditional trials alike, by ensuring that intervention and control groups are broadly similar on certain key variables. Furthermore, such allocation schemes typically require analytic models that account for this clustering and the resulting correlations among error structures (e.g., generalized estimating equations [GEE] or mixed-effects models; Schildcrout et al., 2018 ).

1.1. Optimization trials

Key research questions in implementation science often involve determining which implementation strategies to provide, to whom, and when, to achieve optimal implementation success. As such, trials designed to evaluate comparative effectiveness, or to optimize provision of different types or intensities of implementation strategies, may be more appealing than traditional effectiveness trials. The methods described in this section are not unique to implementation science, but their application in the context of implementation trials may be particularly useful for informing implementation strategies.

While two-arm RCTs can be used to evaluate comparative effectiveness, trials focused on optimizing implementation support may use alternative experimental designs ( Collins et al., 2005 ; Collins et al., 2007 ). For example, in certain clinical contexts, multi-component “bundles” of implementation strategies may be warranted (e.g. a bundle consisting of clinician training, technical assistance, and audit/feedback to encourage clinicians to use a new evidence-based practice). In these situations, implementation researchers might consider using factorial or fractional-factorial designs. In the context of implementation science, these designs randomize participants (e.g. sites or providers) to different combinations of implementation strategies, and can be used to evaluate the effectiveness of each strategy individually to inform an optimal combination (e.g. Coulton et al., 2009 ; Pellegrini et al., 2014 ; Wyrick, et al., 2014 ). Such designs can be particularly useful in informing multi-component implementation strategies that are not redundant or overly burdensome ( Collins et al., 2014a ; Collins et al., 2009 ; Collins et al., 2007 ).

Researchers interested in optimizing sequences of implementation strategies that adapt to ongoing needs over time may be interested in a variant of factorial designs known as the sequential, multiple-assignment randomized trial (SMART; Almirall et al., 2012 ; Collins et al., 2014b ; Kilbourne et al., 2014b ; Lei et al., 2012 ; Nahum-Shani et al., 2012 ; NeCamp et al., 2017 ). SMARTs are multistage randomized trials in which some or all participants are randomized more than once, often based on ongoing information (e.g., treatment response). In implementation research, SMARTs can inform optimal sequences of implementation strategies to maximize downstream clinical outcomes. Thus, such designs are well-suited to answering questions about what implementation strategies should be used, in what order, to achieve the best outcomes in a given context.

One example of an implementation SMART is the Adaptive Implementation of Effective Program Trial (ADEPT; Kilbourne et al., 2014a ). ADEPT was a clustered SMART ( NeCamp et al., 2017 ) designed to inform an adaptive sequence of implementation strategies for implementing an evidence-based collaborative chronic care model, Life Goals ( Kilbourne et al., 2014c ; Kilbourne et al., 2012a ), into community-based practices. Life Goals, the clinical intervention being implemented, has proven effective at improving physical and mental health outcomes for patients with unipolar and bipolar depression by encouraging providers to instruct patients in self-management, and improving clinical information systems and care management across physical and mental health providers ( Bauer et al., 2006 ; Kilbourne et al., 2012a ; Kilbourne et al., 2008 ; Simon et al., 2006 ). However, in spite of its established clinical effectiveness, community-based clinics experienced a number of barriers in trying to implement the Life Goals model, and there were questions about how best to efficiently and effectively augment implementation strategies for clinics that struggled with implementation.

The ADEPT study was thus designed to determine the best sequence of implementation strategies to offer sites interested in implementing Life Goals. The ADEPT study involved use of three different implementation strategies. First, all sites received implementation support based on Replicating Effective Programs (REP), which offered an implementation manual, brief training, and low- level technical support ( Kilbourne et al., 2007 ; Kilbourne et al., 2012b ; Neumann and Sogolow, 2000 ). REP implementation support had been previously found to be low-cost and readily scalable, but also insufficient for uptake for many community-based settings ( Kilbourne et al., 2015 ). For sites that failed to implement Life Goals under REP, two additional implementation strategies were considered as augmentations to REP: External Facilitation (EF; Kilbourne et al., 2014b ; Stetler et al., 2006 ), consisting of phone-based mentoring in strategic skills from a study team member; and Internal Facilitation (IF; Kirchner et al., 2014 ), which supported protected time for a site employee to address barriers to program adoption.

The ADEPT study was designed to evaluate the best way to augment support for these sites that were not able to implement Life Goals under REP, specifically querying whether it was better to augment REP with EF only or the more intensive EF/IF, and whether augmentations should be provided all at once, or staged. Intervention assignments are mapped in Figure 1 . Seventy-nine community-based clinics across Michigan and Colorado were provided with initial implementation support under REP. After six months, implementation of the clinical intervention, Life Goals, was evaluated at all sites. Sites that had failed to reach an adequate level of delivery (defined as those sites enrolling fewer than ten patients in Life Goals, or those at which fewer than 50% of enrolled patients had received at least three Life Goals sessions) were considered non-responsive to REP and randomized to receive additional support through either EF or combined EF/IF. After six further months, Life Goals implementation at these sites was again evaluated. Sites surpassing the implementation response benchmark had their EF or EF/IF support discontinued. EF/IF sites that remained non-responsive continued to receive EF/IF for an additional six months. EF sites that remained non-responsive were randomized a second time to either continue with EF or further augment with IF. This design thus allowed for comparison of three different adaptive implementation interventions for sites that were initially non-responsive to REP to determine the best adaptive sequence of implementation support for sites that were initially non-responsive under REP:

An external file that holds a picture, illustration, etc.
Object name is nihms-1533574-f0001.jpg

SMART design from ADEPT trial.

  • Provide EF for 6 months; continue EF for a further six months for sites that remain nonresponsive; discontinue EF for sites that are responsive;
  • Provide EF/IF for 6 months; continue EF/IF for a further six months for sites that remain non-responsive; discontinue EF/IF for sites that are responsive; and
  • Provide EF for 6 months; step up to EF/IF for a further six months for sites that remain non-responsive; discontinue EF for sites that are responsive.

While analyses of this study are still ongoing, including the comparison of these three adaptive sequences of implementation strategies, results have shown that patients at sites that were randomized to receive EF as the initial augmentation to REP saw more improvement in clinical outcomes (SF-12 mental health quality of life and PHQ-9 depression scores) after 12 months than patients at sites that were randomized to receive the more intensive EF/IF augmentation.

1.2. Single Subject Experimental Designs and On-Off-On (ABA) Designs

We also note that there are a variety of Single Subject Experimental Designs (SSEDs; Byiers et al., 2012 ), including withdrawal designs and alternating treatment designs, that can be used in testing evidence-based practices. Similarly, an implementation strategy may be used to encourage the use of a specific treatment at a particular site, followed by that strategy’s withdrawal and subsequent reinstatement, with data collection throughout the process (on-off-on or ABA design). A weakness of these approaches in the context of implementation science, however, is that they usually require reversibility of the intervention (i.e. that the withdrawal of implementation support truly allows the healthcare system to revert to its pre-implementation state). When this is not the case—for example, if a hypothetical study is focused on training to encourage use of an evidence-based psychotherapy—then these designs may be less useful.

2. Quasi-Experimental Designs in Implementation Science

In some implementation science contexts, policy-makers or administrators may not be willing to have a subset of participating patients or sites randomized to a control condition, especially for high-profile or high-urgency clinical issues. Quasi-experimental designs allow implementation scientists to conduct rigorous studies in these contexts, albeit with certain limitations. We briefly review the characteristics of these designs here; other recent review articles are available for the interested reader (e.g. Handley et al., 2018 ).

2.1. Pre-Post with Non-Equivalent Control Group

The pre-post with non-equivalent control group uses a control group in the absence of randomization. Ideally, the control group is chosen to be as similar to the intervention group as possible (e.g. by matching on factors such as clinic type, patient population, geographic region, etc.). Theoretically, both groups are exposed to the same trends in the environment, making it plausible to decipher if the intervention had an effect. Measurement of both treatment and control conditions classically occurs pre- and post-intervention, with differential improvement between the groups attributed to the intervention. This design is popular due to its practicality, especially if data collection points can be kept to a minimum. It may be especially useful for capitalizing on naturally occurring experiments such as may occur in the context of certain policy initiatives or rollouts—specifically, rollouts in which it is plausible that a control group can be identified. For example, Kirchner and colleagues (2014) used this type of design to evaluate the integration of mental health services into primary care clinics at seven US Department of Veterans Affairs (VA) medical centers and seven matched controls.

One overarching drawback of this design is that it is especially vulnerable to threats to internal validity ( Shadish, 2002 ), because pre-existing differences between the treatment and control group could erroneously be attributed to the intervention. While unmeasured differences between treatment and control groups are always a possibility in healthcare research, such differences are especially likely to occur in the context of these designs due to the lack of randomization. Similarly, this design is particularly sensitive to secular trends that may differentially affect the treatment and control groups ( Cousins et al., 2014 ; Pape et al., 2013 ), as well as regression to the mean confounding study results ( Morton and Torgerson, 2003 ). For example, if a study site is selected for the experimental condition precisely because it is underperforming in some way, then regression to the mean would suggest that the site will show improvement regardless of any intervention; in the context of a pre-post with non-equivalent control group study, however, this improvement would erroneously be attributed to the intervention itself (Type I error).

There are, however, various ways that implementation scientists can mitigate these weaknesses. First, as mentioned briefly above, it is important to select a control group that is as similar as possible to the intervention site(s), which can include matching at both the health care network and clinic level (e.g. Kirchner et al., 2014 ). Second, propensity score weighting (e.g. Morgan, 2018 ) can statistically mitigate internal validity concerns, although this approach may be of limited utility when comparing secular trends between different study cohorts ( Dimick and Ryan, 2014 ). More broadly, qualitative methods (e.g. periodic interviews with staff at intervention and control sites) can help uncover key contextual factors that may be affecting study results above and beyond the intervention itself.

2.2. Interrupted Time Series

Interrupted time series (ITS; Shadish, 2002 ; Taljaard et al., 2014 ; Wagner et al., 2002 ) designs represent one of the most robust categories of quasi-experimental designs. Rather than relying on a non-equivalent control group, ITS designs rely on repeated data collections from intervention sites to determine whether a particular intervention is associated with improvement on a given metric relative to the pre-intervention secular trend. They are particularly useful in cases where a comparable control group cannot be identified—for example, following widespread implementation of policy mandates, quality improvement initiatives, or dissemination campaigns ( Eccles et al., 2003 ). In ITS designs, data are collected at multiple time points both before and after an intervention (e.g., policy change, implementation effort), and analyses explore whether the intervention was associated with the outcome beyond any pre-existing secular trend. More formally, ITS evaluations focus on identifying whether there is discontinuity in the trend (change in slope or level) after the intervention relative to before the intervention, using segmented regression to model pre- and post-intervention trends ( Gebski et al., 2012 ; Penfold and Zhang, 2013 ; Taljaard et al., 2014 ; Wagner et al., 2002 ). A number of recent implementation studies have used ITS designs, including an evaluation of implementation of a comprehensive smoke-free policy in a large UK mental health organization to reduce physical assaults ( Robson et al., 2017 ); the impact of a national policy limiting alcohol availability on suicide mortality in Slovenia ( Pridemore and Snowden, 2009 ); and the effect of delivery of a tailored intervention for primary care providers to increase psychological referrals for women with mild to moderate postnatal depression ( Hanbury et al., 2013 ).

ITS designs are appealing in implementation work for several reasons. Relative to uncontrolled pre-post analyses, ITS analyses reduce the chances that intervention effects are confounded by secular trends ( Bernal et al., 2017 ; Eccles et al., 2003 ). Time-varying confounders, such as seasonality, can also be adjusted for, provided adequate data ( Bernal et al., 2017 ). Indeed, recent work has confirmed that ITS designs can yield effect estimates similar to those derived from cluster-randomized RCTs ( Fretheim et al., 2013 ; Fretheim et al., 2015 ). Relative to an RCT, ITS designs can also allow for a more comprehensive assessment of the longitudinal effects of an intervention (positive or negative), as effects can be traced over all included time points ( Bernal et al., 2017 ; Penfold and Zhang, 2013 ).

ITS designs also present a number of challenges. First, the segmented regression approach requires clear delineation between pre- and post-intervention periods; interventions with indeterminate implementation periods are likely not good candidates for ITS. While ITS designs that include multiple ‘interruptions’ (e.g. introductions of new treatment components) are possible, they will require collection of enough time points between interruptions to ensure that each intervention’s effects can be ascertained individually ( Bernal et al., 2017 ). Second, collecting data from sufficient time points across all sites of interest, especially for the pre-intervention period, can be challenging ( Eccles et al., 2003 ): a common recommendation is at least eight time points both pre- and post-intervention ( Penfold and Zhang, 2013 ). This may be onerous, particularly if the data are not routinely collected by the health system(s) under study. Third, ITS cannot protect against confounding effects from other interventions that begin contemporaneously and may impact similar outcomes ( Eccles et al., 2003 ).

2.3. Stepped Wedge Designs

Stepped wedge trials are another type of quasi-experimental design. In a stepped wedge, all participants receive the intervention, but are assigned to the timing of the intervention in a staggered fashion ( Betran et al., 2018 ; Brown and Lilford, 2006 ; Hussey and Hughes, 2007 ), typically at the site or cluster level. Stepped wedge designs have their analytic roots in balanced incomplete block designs, in which all pairs of treatments occur an equal number of times within each block ( Hanani, 1961 ). Traditionally, all sites in stepped wedge trials have outcome measures assessed at all time points, thus allowing sites that receive the intervention later in the trial to essentially serve as controls for early intervention sites. A recent special issue of the journal Trials includes more detail on these designs ( Davey et al., 2015 ), which may be ideal for situations in which it is important for all participating patients or sites to receive the intervention during the trial. Stepped wedge trials may also be useful when resources are scarce enough that intervening at all sites at once (or even half of the sites as in a standard treatment-versus-control RCT) would not be feasible. If desired, the administration of the intervention to sites in waves allows for lessons learned in early sites to be applied to later sites (via formative evaluation; see Elwy et al., this issue).

The Behavioral Health Interdisciplinary Program (BHIP) Enhancement Project is a recent example of a stepped-wedge implementation trial ( Bauer et al., 2016 ; Bauer et al., 2019 ). This study involved using blended facilitation (including internal and external facilitators; Kirchner et al., 2014 ) to implement care practices consistent with the collaborative chronic care model (CCM; Bodenheimer et al., 2002a , b ; Wagner et al., 1996 ) in nine outpatient mental health teams in VA medical centers. Figure 2 illustrates the implementation and stepdown periods for that trial, with black dots representing primary data collection points.

An external file that holds a picture, illustration, etc.
Object name is nihms-1533574-f0002.jpg

BHIP Enhancement Project stepped wedge (adapted form Bauer et al., 2019).

The BHIP Enhancement Project was conducted as a stepped wedge for several reasons. First, the stepped wedge design allowed the trial to reach nine sites despite limited implementation resources (i.e. intervening at all nine sites simultaneously would not have been feasible given study funding). Second, the stepped wedge design aided in recruitment and retention, as all participating sites were certain to receive implementation support during the trial: at worst, sites that were randomized to later- phase implementation had to endure waiting periods totaling about eight months before implementation began. This was seen as a major strength of the design by its operational partner, the VA Office of Mental Health and Suicide Prevention. To keep sites engaged during the waiting period, the BHIP Enhancement Project offered a guiding workbook and monthly technical support conference calls.

Three additional features of the BHIP Enhancement Project deserve special attention. First, data collection for late-implementing sites did not begin until immediately before the onset of implementation support (see Figure 2 ). While this reduced statistical power, it also significantly reduced data collection burden on the study team. Second, onset of implementation support was staggered such that wave 2 began at the end of month 4 rather than month 6. This had two benefits: first, this compressed the overall amount of time required for implementation during the trial. Second, it meant that the study team only had to collect data from one site at a time, with data collection periods coming every 2–4 months. More traditional stepped wedge approaches typically have data collection across sites temporally aligned (e.g. Betran et al., 2018 ). Third, the BHIP Enhancement Project used a balancing algorithm ( Lew et al., 2019 ) to assign sites to waves, retaining some of the benefits of randomization while ensuring balance on key site characteristics (e.g. size, geographic region).

Despite their utility, stepped wedges have some important limitations. First, because they feature delayed implementation at some sites, stepped wedges typically take longer than similarly-sized parallel group RCTs. This increases the chances that secular trends, policy changes, or other external forces impact study results. Second, as with RCTs, imbalanced site assignment can confound results. This may occur deliberately in some cases—for example, if sites that develop their implementation plans first are assigned to earlier waves. Even if sites are randomized, however, early and late wave sites may still differ on important characteristics such as size, rurality, and case mix. The resulting confounding between site assignment and time can threaten the internal validity of the study—although, as above, balancing algorithms can reduce this risk. Third, the use of formative evaluation (Elwy, this issue), while useful for maximizing the utility of implementation efforts in a stepped wedge, can mean that late-wave sites receive different implementation strategies than early-wave sites. Similarly, formative evaluation may inform midstream adaptations to the clinical innovation being implemented. In either case, these changes may again threaten internal validity. Overall, then, stepped wedges represent useful tools for evaluating the impact of health interventions that (as with all designs) are subject to certain weaknesses and limitations.

3. Conclusions and Future Directions

Implementation science is focused on maximizing the extent to which effective healthcare practices are adopted, used, and sustained by clinicians, hospitals, and systems. Answering questions in these domains frequently requires different research methods than those employed in traditional efficacy- or effectiveness-oriented randomized clinical trials (RCTs). Implementation-oriented RCTs typically feature cluster or site-level randomization, and emphasize implementation outcomes (e.g. the number of patients receiving the new treatment as intended) rather than traditional clinical outcomes. Hybrid implementation-effectiveness designs incorporate both types of outcomes; more details on these approaches can be found elsewhere in this special issue (Landes, this issue). Other methodological innovations, such as factorial designs or sequential, multiple-assignment randomized trials (SMARTs), can address questions about multi-component or adaptive interventions, still under the umbrella of experimental designs. These types of trials may be especially important for demystifying the “black box” of implementation—that is, determining what components of an implementation strategy are most strongly associated with implementation success. In contrast, pre-post designs with non-equivalent control groups, interrupted time series (ITS), and stepped wedge designs are all examples of quasiexperimental designs that may serve implementation researchers when experimental designs would be inappropriate. A major theme cutting across each of these designs is that there are relative strengths and weaknesses associated with any study design decision. Determining what design to use ultimately will need to be informed by the primary research question to be answered, while simultaneously balancing the need for internal validity, external validity, feasibility, and ethics.

New innovations in study design are constantly being developed and refined. Several such innovations are covered in other articles within this special issue (e.g. Kim et al., this issue). One future direction relevant to the study designs presented in this article is the potential for adaptive trial designs, which allow information gleaned during the trial to inform the adaptation of components like treatment allocation, sample size, or study recruitment in the later phases of the same trial ( Pallmann et al., 2018 ). These designs are becoming increasingly popular in clinical treatment ( Bhatt and Mehta, 2016 ) but could also hold promise for implementation scientists, especially as interest grows in rapid-cycle testing of implementation strategies or efforts. Adaptive designs could potentially be incorporated into both SMART designs and stepped wedge studies, as well as traditional RCTs to further advance implementation science ( Cheung et al., 2015 ). Ideally, these and other innovations will provide researchers with increasingly robust and useful methodologies for answering timely implementation science questions.

  • Many implementation science questions can be addressed by fully experimental designs (e.g. randomized controlled trials [RCTs]).
  • Implementation trials differ in important ways, however, from more traditional efficacy- or effectiveness-oriented RCTs.
  • Adaptive designs represent a recent innovation to determine optimal implementation strategies within a fully experimental framework.
  • Quasi-experimental designs can be used to answer implementation science questions in the absence of randomization.
  • The choice of study designs in implementation science requires careful consideration of scientific, pragmatic, and ethical issues.

Acknowledgments

This work was supported by Department of Veterans Affairs grants QUE 15–289 (PI: Bauer) and CIN 13403 and National Institutes of Health grant RO1 MH 099898 (PI: Kilbourne).

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

  • Almirall D, Compton SN, Gunlicks-Stoessel M, Duan N, Murphy SA, 2012. Designing a pilot sequential multiple assignment randomized trial for developing an adaptive treatment strategy . Stat Med 31 ( 17 ), 1887–1902. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bauer MS, McBride L, Williford WO, Glick H, Kinosian B, Altshuler L, Beresford T, Kilbourne AM, Sajatovic M, Cooperative Studies Program 430 Study, T., 2006. Collaborative care for bipolar disorder: Part II. Impact on clinical outcome, function, and costs . Psychiatr Serv 57 ( 7 ), 937–945. [ PubMed ] [ Google Scholar ]
  • Bauer MS, Miller C, Kim B, Lew R, Weaver K, Coldwell C, Henderson K, Holmes S, Seibert MN, Stolzmann K, Elwy AR, Kirchner J, 2016. Partnering with health system operations leadership to develop a controlled implementation trial . Implement Sci 11 , 22. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bauer MS, Miller CJ, Kim B, Lew R, Stolzmann K, Sullivan J, Riendeau R, Pitcock J, Williamson A, Connolly S, Elwy AR, Weaver K, 2019. Effectiveness of Implementing a Collaborative Chronic Care Model for Clinician Teams on Patient Outcomes and Health Status in Mental Health: A Randomized Clinical Trial . JAMA Netw Open 2 ( 3 ), e190230. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bernal JL, Cummins S, Gasparrini A, 2017. Interrupted time series regression for the evaluation of public health interventions: a tutorial . Int J Epidemiol 46 ( 1 ), 348–355. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Betran AP, Bergel E, Griffin S, Melo A, Nguyen MH, Carbonell A, Mondlane S, Merialdi M, Temmerman M, Gulmezoglu AM, 2018. Provision of medical supply kits to improve quality of antenatal care in Mozambique: a stepped-wedge cluster randomised trial . Lancet Glob Health 6 ( 1 ), e57–e65. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Bhatt DL, Mehta C, 2016. Adaptive Designs for Clinical Trials . N Engl J Med 375 ( 1 ), 65–74. [ PubMed ] [ Google Scholar ]
  • Bodenheimer T, Wagner EH, Grumbach K, 2002a. Improving primary care for patients with chronic illness . JAMA 288 ( 14 ), 1775–1779. [ PubMed ] [ Google Scholar ]
  • Bodenheimer T, Wagner EH, Grumbach K, 2002b. Improving primary care for patients with chronic illness: the chronic care model, Part 2 . JAMA 288 ( 15 ), 1909–1914. [ PubMed ] [ Google Scholar ]
  • Brown CA, Lilford RJ, 2006. The stepped wedge trial design: a systematic review . BMC medical research methodology 6 ( 1 ), 54. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Byiers BJ, Reichle J, Symons FJ, 2012. Single-subject experimental design for evidence-based practice . Am J Speech Lang Pathol 21 ( 4 ), 397–414. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cheung YK, Chakraborty B, Davidson KW, 2015. Sequential multiple assignment randomized trial (SMART) with adaptive randomization for quality improvement in depression treatment program . Biometrics 71 ( 2 ), 450–459. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Collins LM, Dziak JJ, Kugler KC, Trail JB, 2014a. Factorial experiments: efficient tools for evaluation of intervention components . Am J Prev Med 47 ( 4 ), 498–504. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Collins LM, Dziak JJ, Li R, 2009. Design of experiments with multiple independent variables: a resource management perspective on complete and reduced factorial designs . Psychol Methods 14 ( 3 ), 202–224. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Collins LM, Murphy SA, Bierman KL, 2004. A conceptual framework for adaptive preventive interventions . Prev Sci 5 ( 3 ), 185–196. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Collins LM, Murphy SA, Nair VN, Strecher VJ, 2005. A strategy for optimizing and evaluating behavioral interventions . Ann Behav Med 30 ( 1 ), 65–73. [ PubMed ] [ Google Scholar ]
  • Collins LM, Murphy SA, Strecher V, 2007. The multiphase optimization strategy (MOST) and the sequential multiple assignment randomized trial (SMART): new methods for more potent eHealth interventions . Am J Prev Med 32 ( 5 Suppl ), S112–118. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Collins LM, Nahum-Shani I, Almirall D, 2014b. Optimization of behavioral dynamic treatment regimens based on the sequential, multiple assignment, randomized trial (SMART) . Clin Trials 11 ( 4 ), 426–434. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Coulton S, Perryman K, Bland M, Cassidy P, Crawford M, Deluca P, Drummond C, Gilvarry E, Godfrey C, Heather N, Kaner E, Myles J, Newbury-Birch D, Oyefeso A, Parrott S, Phillips T, Shenker D, Shepherd J, 2009. Screening and brief interventions for hazardous alcohol use in accident and emergency departments: a randomised controlled trial protocol . BMC Health Serv Res 9 , 114. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Cousins K, Connor JL, Kypri K, 2014. Effects of the Campus Watch intervention on alcohol consumption and related harm in a university population . Drug Alcohol Depend 143 , 120–126. [ PubMed ] [ Google Scholar ]
  • Curran GM, Bauer M, Mittman B, Pyne JM, Stetler C, 2012. Effectiveness-implementation hybrid designs: combining elements of clinical effectiveness and implementation research to enhance public health impact . Med Care 50 ( 3 ), 217–226. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Davey C, Hargreaves J, Thompson JA, Copas AJ, Beard E, Lewis JJ, Fielding KL, 2015. Analysis and reporting of stepped wedge randomised controlled trials: synthesis and critical appraisal of published studies, 2010 to 2014 . Trials 16 ( 1 ), 358. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Dimick JB, Ryan AM, 2014. Methods for evaluating changes in health care policy: the difference-in- differences approach . JAMA 312 ( 22 ), 2401–2402. [ PubMed ] [ Google Scholar ]
  • Eccles M, Grimshaw J, Campbell M, Ramsay C, 2003. Research designs for studies evaluating the effectiveness of change and improvement strategies . Qual Saf Health Care 12 ( 1 ), 47–52. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Fisher RA, 1925, July Theory of statistical estimation In Mathematical Proceedings of the Cambridge Philosophical Society (Vol. 22, No. 5, pp. 700–725). Cambridge University Press. [ Google Scholar ]
  • Fisher RA, 1935. The design of experiments . Oliver and Boyd, Edinburgh. [ Google Scholar ]
  • Fretheim A, Soumerai SB, Zhang F, Oxman AD, Ross-Degnan D, 2013. Interrupted time-series analysis yielded an effect estimate concordant with the cluster-randomized controlled trial result . Journal of Clinical Epidemiology 66 ( 8 ), 883–887. [ PubMed ] [ Google Scholar ]
  • Fretheim A, Zhang F, Ross-Degnan D, Oxman AD, Cheyne H, Foy R, Goodacre S, Herrin J, Kerse N, McKinlay RJ, Wright A, Soumerai SB, 2015. A reanalysis of cluster randomized trials showed interrupted time-series studies were valuable in health system evaluation . J Clin Epidemiol 68 ( 3 ), 324–333. [ PubMed ] [ Google Scholar ]
  • Gaglio B, Shoup JA, Glasgow RE, 2013. The RE-AIM framework: a systematic review of use over time . Am J Public Health 103 ( 6 ), e38–46. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Gebski V, Ellingson K, Edwards J, Jernigan J, Kleinbaum D, 2012. Modelling interrupted time series to evaluate prevention and control of infection in healthcare . Epidemiol Infect 140 ( 12 ), 2131–2141. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Glasgow RE, Vogt TM, Boles SM, 1999. Evaluating the public health impact of health promotion interventions: the RE-AIM framework . Am J Public Health 89 ( 9 ), 1322–1327. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hanani H, 1961. The existence and construction of balanced incomplete block designs . The Annals of Mathematical Statistics 32 ( 2 ), 361–386. [ Google Scholar ]
  • Hanbury A, Farley K, Thompson C, Wilson PM, Chambers D, Holmes H, 2013. Immediate versus sustained effects: interrupted time series analysis of a tailored intervention . Implement Sci 8 , 130. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Handley MA, Lyles CR, McCulloch C, Cattamanchi A, 2018. Selecting and Improving Quasi-Experimental Designs in Effectiveness and Implementation Research . Annu Rev Public Health 39 , 5–25. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hussey MA, Hughes JP, 2007. Design and analysis of stepped wedge cluster randomized trials . Contemp Clin Trials 28 ( 2 ), 182–191. [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Almirall D, Eisenberg D, Waxmonsky J, Goodrich DE, Fortney JC, Kirchner JE, Solberg LI, Main D, Bauer MS, Kyle J, Murphy SA, Nord KM, Thomas MR, 2014a. Protocol: Adaptive Implementation of Effective Programs Trial (ADEPT): cluster randomized SMART trial comparing a standard versus enhanced implementation strategy to improve outcomes of a mood disorders program . Implement Sci 9 , 132. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Almirall D, Goodrich DE, Lai Z, Abraham KM, Nord KM, Bowersox NW, 2014b. Enhancing outreach for persons with serious mental illness: 12-month results from a cluster randomized trial of an adaptive implementation strategy . Implement Sci 9 , 163. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Bramlet M, Barbaresso MM, Nord KM, Goodrich DE, Lai Z, Post EP, Almirall D, Verchinina L, Duffy SA, Bauer MS, 2014c. SMI life goals: description of a randomized trial of a collaborative care model to improve outcomes for persons with serious mental illness . Contemp Clin Trials 39 ( 1 ), 74–85. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Goodrich DE, Lai Z, Clogston J, Waxmonsky J, Bauer MS, 2012a. Life Goals Collaborative Care for patients with bipolar disorder and cardiovascular disease risk . Psychiatr Serv 63 ( 12 ), 1234–1238. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Goodrich DE, Nord KM, Van Poppelen C, Kyle J, Bauer MS, Waxmonsky JA, Lai Z, Kim HM, Eisenberg D, Thomas MR, 2015. Long-Term Clinical Outcomes from a Randomized Controlled Trial of Two Implementation Strategies to Promote Collaborative Care Attendance in Community Practices . Adm Policy Ment Health 42 ( 5 ), 642–653. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Neumann MS, Pincus HA, Bauer MS, Stall R, 2007. Implementing evidence-based interventions in health care: application of the replicating effective programs framework . Implement Sci 2 , 42. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Neumann MS, Waxmonsky J, Bauer MS, Kim HM, Pincus HA, Thomas M, 2012b. Public-academic partnerships: evidence-based implementation: the role of sustained community-based practice and research partnerships . Psychiatr Serv 63 ( 3 ), 205–207. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Kilbourne AM, Post EP, Nossek A, Drill L, Cooley S, Bauer MS, 2008. Improving medical and psychiatric outcomes among individuals with bipolar disorder: a randomized controlled trial . Psychiatr Serv 59 ( 7 ), 760–768. [ PubMed ] [ Google Scholar ]
  • Kirchner JE, Ritchie MJ, Pitcock JA, Parker LE, Curran GM, Fortney JC, 2014. Outcomes of a partnered facilitation strategy to implement primary care-mental health . J Gen Intern Med 29 Suppl 4 , 904–912. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Lei H, Nahum-Shani I, Lynch K, Oslin D, Murphy SA, 2012. A “SMART” design for building individualized treatment sequences . Annu Rev Clin Psychol 8 , 21–48. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Lew RA, Miller CJ, Kim B, Wu H, Stolzmann K, Bauer MS, 2019. A robust method to reduce imbalance for site-level randomized controlled implementation trial designs . Implementation Sci , 14 , 46. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Morgan CJ, 2018. Reducing bias using propensity score matching . J Nucl Cardiol 25 ( 2 ), 404–406. [ PubMed ] [ Google Scholar ]
  • Morton V, Torgerson DJ, 2003. Effect of regression to the mean on decision making in health care . BMJ 326 ( 7398 ), 1083–1084. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Nahum-Shani I, Qian M, Almirall D, Pelham WE, Gnagy B, Fabiano GA, Waxmonsky JG, Yu J, Murphy SA, 2012. Experimental design and primary data analysis methods for comparing adaptive interventions . Psychol Methods 17 ( 4 ), 457–477. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • NeCamp T, Kilbourne A, Almirall D, 2017. Comparing cluster-level dynamic treatment regimens using sequential, multiple assignment, randomized trials: Regression estimation and sample size considerations . Stat Methods Med Res 26 ( 4 ), 1572–1589. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Neumann MS, Sogolow ED, 2000. Replicating effective programs: HIV/AIDS prevention technology transfer . AIDS Educ Prev 12 ( 5 Suppl ), 35–48. [ PubMed ] [ Google Scholar ]
  • Pallmann P, Bedding AW, Choodari-Oskooei B, Dimairo M, Flight L, Hampson LV, Holmes J, Mander AP, Odondi L.o., Sydes MR, Villar SS, Wason JMS, Weir CJ, Wheeler GM, Yap C, Jaki T, 2018. Adaptive designs in clinical trials: why use them, and how to run and report them . BMC medicine 16 ( 1 ), 29–29. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Pape UJ, Millett C, Lee JT, Car J, Majeed A, 2013. Disentangling secular trends and policy impacts in health studies: use of interrupted time series analysis . J R Soc Med 106 ( 4 ), 124–129. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Pellegrini CA, Hoffman SA, Collins LM, Spring B, 2014. Optimization of remotely delivered intensive lifestyle treatment for obesity using the Multiphase Optimization Strategy: Opt-IN study protocol . Contemp Clin Trials 38 ( 2 ), 251–259. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Penfold RB, Zhang F, 2013. Use of Interrupted Time Series Analysis in Evaluating Health Care Quality Improvements . Academic Pediatrics 13 ( 6, Supplement ), S38–S44. [ PubMed ] [ Google Scholar ]
  • Pridemore WA, Snowden AJ, 2009. Reduction in suicide mortality following a new national alcohol policy in Slovenia: an interrupted time-series analysis . Am J Public Health 99 ( 5 ), 915–920. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Proctor E, Silmere H, Raghavan R, Hovmand P, Aarons G, Bunger A, Griffey R, Hensley M, 2011. Outcomes for implementation research: conceptual distinctions, measurement challenges, and research agenda . Adm Policy Ment Health 38 ( 2 ), 65–76. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Robson D, Spaducci G, McNeill A, Stewart D, Craig TJK, Yates M, Szatkowski L, 2017. Effect of implementation of a smoke-free policy on physical violence in a psychiatric inpatient setting: an interrupted time series analysis . Lancet Psychiatry 4 ( 7 ), 540–546. [ PubMed ] [ Google Scholar ]
  • Schildcrout JS, Schisterman EF, Mercaldo ND, Rathouz PJ, Heagerty PJ, 2018. Extending the Case-Control Design to Longitudinal Data: Stratified Sampling Based on Repeated Binary Outcomes . Epidemiology 29 ( 1 ), 67–75. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Shadish WR, Cook Thomas D., Campbell Donald T, 2002. Experimental and quasi-experimental designs for generalized causal inference . Houghton Miffflin Company, Boston, MA. [ Google Scholar ]
  • Simon GE, Ludman EJ, Bauer MS, Unutzer J, Operskalski B, 2006. Long-term effectiveness and cost of a systematic care program for bipolar disorder . Arch Gen Psychiatry 63 ( 5 ), 500–508. [ PubMed ] [ Google Scholar ]
  • Stetler CB, Legro MW, Rycroft-Malone J, Bowman C, Curran G, Guihan M, Hagedorn H, Pineros S, Wallace CM, 2006. Role of “external facilitation” in implementation of research findings: a qualitative evaluation of facilitation experiences in the Veterans Health Administration . Implement Sci 1 , 23. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Taljaard M, McKenzie JE, Ramsay CR, Grimshaw JM, 2014. The use of segmented regression in analysing interrupted time series studies: an example in pre-hospital ambulance care . Implement Sci 9 , 77. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Wagner AK, Soumerai SB, Zhang F, Ross-Degnan D, 2002. Segmented regression analysis of interrupted time series studies in medication use research . J Clin Pharm Ther 27 ( 4 ), 299–309. [ PubMed ] [ Google Scholar ]
  • Wagner EH, Austin BT, Von Korff M, 1996. Organizing care for patients with chronic illness . Milbank Q 74 ( 4 ), 511–544. [ PubMed ] [ Google Scholar ]
  • Wyrick DL, Rulison KL, Fearnow-Kenney M, Milroy JJ, Collins LM, 2014. Moving beyond the treatment package approach to developing behavioral interventions: addressing questions that arose during an application of the Multiphase Optimization Strategy (MOST) . Transl Behav Med 4 ( 3 ), 252–259. [ PMC free article ] [ PubMed ] [ Google Scholar ]

IMAGES

  1. Chapter 5: Experimental and Quasi-Experimental Designs

    random assignment quasi experimental

  2. Random Assignment Is Used in Experiments Because Researchers Want to

    random assignment quasi experimental

  3. true vs quasi-experimental research: the hallmark difference being that

    random assignment quasi experimental

  4. PPT

    random assignment quasi experimental

  5. 5 Quasi-Experimental Design Examples (2024)

    random assignment quasi experimental

  6. PPT

    random assignment quasi experimental

VIDEO

  1. AP Stats 4.2D Video

  2. quasi-experimental design # M.ed. assignment # ms University saharanpur

  3. random sampling & assignment

  4. Experimental Design AP Practice Question #1

  5. QUASI

  6. 0-Coast Patch of the Week #17: Quasi Random Gates

COMMENTS

  1. Quasi-Experimental Design

    Revised on January 22, 2024. Like a true experiment, a quasi-experimental design aims to establish a cause-and-effect relationship between an independent and dependent variable. However, unlike a true experiment, a quasi-experiment does not rely on random assignment. Instead, subjects are assigned to groups based on non-random criteria.

  2. Quasi Experimental Design Overview & Examples

    In contrast, true experiments use random assignment to the treatment and control groups to control confounding variables, making them the gold standard for identifying cause-and-effect relationships.. Quasi-experimental research is a design that closely resembles experimental research but is different. The term "quasi" means "resembling," so you can think of it as a cousin to actual ...

  3. Quasi-experiment

    A quasi-experiment is an empirical interventional study used to estimate the causal impact of an intervention on target population without random assignment.Quasi-experimental research shares similarities with the traditional experimental design or randomized controlled trial, but it specifically lacks the element of random assignment to treatment or control.

  4. Random Assignment in Experiments

    In experimental research, random assignment is a way of placing participants from your sample into different treatment groups using randomization. With simple random assignment, ... When you can't assign participants to groups, you can also conduct a quasi-experimental study. In a quasi-experiment, you study the outcomes of pre-existing ...

  5. 7.3 Quasi-Experimental Research

    Key Takeaways. Quasi-experimental research involves the manipulation of an independent variable without the random assignment of participants to conditions or orders of conditions. Among the important types are nonequivalent groups designs, pretest-posttest, and interrupted time-series designs.

  6. Selecting and Improving Quasi-Experimental Designs in Effectiveness and

    RCTs can also involve random assignment of groups (e.g., clinics, worksites or communities) to intervention and control arms, but a large number of groups are required in order to realize the full benefits of randomization. ... Quasi-experimental designs (QEDs), which first gained prominence in social science research , are increasingly being ...

  7. Quasi-Experimental Designs for Causal Inference

    But under certain conditions quasi-experimental designs that lack random assignment can also be as credible as RCTs (Shadish, Cook, & Campbell, 2002). This article discusses four of the strongest quasi-experimental designs for identifying causal effects: regression discontinuity design, instrumental variable design, matching and propensity ...

  8. Quasi-experimental study designs series—paper 5: a checklist for

    Credible quasi-experimental methods use assignment rules which are either known or can be modeled statistically, including: methods based on a threshold on a continuous scale (or ordinal scale with a minimum number of units) such as a test score (regression discontinuity design) or another form of "exogenous variation" arising, for example ...

  9. Experimental and Quasi-Experimental Methods

    The two key designs that help researchers address whether a program or treatment causes an outcome are the experimental design, which uses random assignment to groups or programs, and quasi-experimental designs, which do not use random assignment (see Shadish et al. 2002; Bell 2010; Trochim 2006). These two methods are important to consider in ...

  10. PDF Quasi-Experimental Designs

    In this reading, we'll discuss five quasi-experimental approaches: 1) matching, 2) mixed designs, 3) single-subject designs, and 4) developmental designs. (b) are plausible causes of the dependent variable. Quasi-experiments are designed to reduce confounding variables as much as possible, given that random assignment is not available.

  11. Quasi-Experimental Design

    Quasi-Experimental Design. Quasi-Experimental Design is a unique research methodology because it is characterized by what is lacks. For example, Abraham & MacDonald (2011) state: ... It differs from experimental research because either there is no control group, no random selection, no random assignment, and/or no active manipulation. ...

  12. Quasi-Experimental Research Design

    No random assignment: Quasi-experimental designs do not require random assignment, which can be difficult or impossible in some cases, such as when studying the effects of a natural disaster. This means that researchers can still make causal inferences, although they must use statistical techniques to control for potential confounding variables

  13. Chapter 8: Quasi-Experimental Research

    The prefix quasi means "resembling." Thus quasi-experimental research is research that resembles experimental research but is not true experimental research. Recall with a true between-groups experiment, random assignment to conditions is used to ensure the groups are equivalent and with a true within-subjects design counterbalancing is used to guard against order effects.

  14. Quasi-Experimental Research

    Key Takeaways. Quasi-experimental research involves the manipulation of an independent variable without the random assignment of participants to conditions or orders of conditions. Among the important types are nonequivalent groups designs, pretest-posttest, and interrupted time-series designs.

  15. 8.2 Quasi-experimental and pre-experimental designs

    Pre-experimental designs - a variation of experimental design that lacks the rigor of experiments and is often used before a true experiment is conducted. Quasi-experimental design - designs lack random assignment to experimental and control groups. Static group design - uses an experimental group and a comparison group, without random ...

  16. 12.2: Pre-experimental and quasi-experimental design

    Quasi-experimental designs are similar to true experiments, but they lack random assignment to experimental and control groups. The most basic of these quasi-experimental designs is the nonequivalent comparison groups design (Rubin & Babbie, 2017). [1] The nonequivalent comparison group design looks a lot like the classic experimental design ...

  17. The Use and Interpretation of Quasi-Experimental Studies in Medical

    Quasi-experimental study designs, often described as nonrandomized, pre-post intervention studies, are common in the medical informatics literature. Yet little has been written about the benefits and limitations of the quasi-experimental approach as applied to informatics studies. ... The lack of random assignment is the major weakness of the ...

  18. (PDF) Experimental and quasi-experimental designs

    Random Assignment: Participants in experimental research are typically randomly assigned to different groups to ensure that each group is similar at the outset of the study.

  19. Experimental vs Quasi-Experimental Design: Which to Choose?

    Random assignment: Non-random assignment (participants get assigned according to their choosing or that of the researcher) Is there a control group? Yes: ... A quasi-experimental design is a non-randomized study design used to evaluate the effect of an intervention. The intervention can be a training program, a policy change or a medical treatment.

  20. Statistical Analysis of Quasi-Experimental Designs:

    Matched Random Assignment. In an experimental design, matched random sampling can be used to equate the groups on one or more characteristics. Whitley (in chapter 8) uses an example of matching on IQ. ... Matching in Quasi-Experimental Designs: Normative Group Matching. Suppose that you have a quasi-experiment where you want to compare an ...

  21. An Introduction to Quasi-Experimental Design

    Quasi-experimental design (QED) is a research design method that's useful when regular experimental conditions are impractical or unethical. ... as it's more difficult to account for all confounding variables without using random assignment The necessary data required for research potentially being inaccurate, outdated, or difficult to ...

  22. Experimental and Quasi-Experimental Designs in Implementation Research

    Other implementation science questions are more suited to quasi-experimental designs, which are intended to estimate the effect of an intervention in the absence of randomization. ... and second, random assignment of subjects. This corresponds to the definition of randomized experiments originally championed by Fisher (1925). From this ...

  23. Quasi-Experimental Design: Types, Examples, Pros, and Cons

    See why leading organizations rely on MasterClass for learning & development. A quasi-experimental design can be a great option when ethical or practical concerns make true experiments impossible, but the research methodology does have its drawbacks. Learn all the ins and outs of a quasi-experimental design.