Have a language expert improve your writing

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

  • Knowledge Base

Methodology

  • Quasi-Experimental Design | Definition, Types & Examples

Quasi-Experimental Design | Definition, Types & Examples

Published on July 31, 2020 by Lauren Thomas . 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.

Quasi-experimental design is a useful tool in situations where true experiments cannot be used for ethical or practical reasons.

Quasi-experimental design vs. experimental design

Table of contents

Differences between quasi-experiments and true experiments, types of quasi-experimental designs, when to use quasi-experimental design, advantages and disadvantages, other interesting articles, frequently asked questions about quasi-experimental designs.

There are several common differences between true and quasi-experimental designs.

True experimental design Quasi-experimental design
Assignment to treatment The researcher subjects to control and treatment groups. Some other, method is used to assign subjects to groups.
Control over treatment The researcher usually . The researcher often , but instead studies pre-existing groups that received different treatments after the fact.
Use of Requires the use of . Control groups are not required (although they are commonly used).

Example of a true experiment vs a quasi-experiment

However, for ethical reasons, the directors of the mental health clinic may not give you permission to randomly assign their patients to treatments. In this case, you cannot run a true experiment.

Instead, you can use a quasi-experimental design.

You can use these pre-existing groups to study the symptom progression of the patients treated with the new therapy versus those receiving the standard course of treatment.

Here's why students love Scribbr's proofreading services

Discover proofreading & editing

Many types of quasi-experimental designs exist. Here we explain three of the most common types: nonequivalent groups design, regression discontinuity, and natural experiments.

Nonequivalent groups design

In nonequivalent group design, the researcher chooses existing groups that appear similar, but where only one of the groups experiences the treatment.

In a true experiment with random assignment , the control and treatment groups are considered equivalent in every way other than the treatment. But in a quasi-experiment where the groups are not random, they may differ in other ways—they are nonequivalent groups .

When using this kind of design, researchers try to account for any confounding variables by controlling for them in their analysis or by choosing groups that are as similar as possible.

This is the most common type of quasi-experimental design.

Regression discontinuity

Many potential treatments that researchers wish to study are designed around an essentially arbitrary cutoff, where those above the threshold receive the treatment and those below it do not.

Near this threshold, the differences between the two groups are often so minimal as to be nearly nonexistent. Therefore, researchers can use individuals just below the threshold as a control group and those just above as a treatment group.

However, since the exact cutoff score is arbitrary, the students near the threshold—those who just barely pass the exam and those who fail by a very small margin—tend to be very similar, with the small differences in their scores mostly due to random chance. You can therefore conclude that any outcome differences must come from the school they attended.

Natural experiments

In both laboratory and field experiments, researchers normally control which group the subjects are assigned to. In a natural experiment, an external event or situation (“nature”) results in the random or random-like assignment of subjects to the treatment group.

Even though some use random assignments, natural experiments are not considered to be true experiments because they are observational in nature.

Although the researchers have no control over the independent variable , they can exploit this event after the fact to study the effect of the treatment.

However, as they could not afford to cover everyone who they deemed eligible for the program, they instead allocated spots in the program based on a random lottery.

Although true experiments have higher internal validity , you might choose to use a quasi-experimental design for ethical or practical reasons.

Sometimes it would be unethical to provide or withhold a treatment on a random basis, so a true experiment is not feasible. In this case, a quasi-experiment can allow you to study the same causal relationship without the ethical issues.

The Oregon Health Study is a good example. It would be unethical to randomly provide some people with health insurance but purposely prevent others from receiving it solely for the purposes of research.

However, since the Oregon government faced financial constraints and decided to provide health insurance via lottery, studying this event after the fact is a much more ethical approach to studying the same problem.

True experimental design may be infeasible to implement or simply too expensive, particularly for researchers without access to large funding streams.

At other times, too much work is involved in recruiting and properly designing an experimental intervention for an adequate number of subjects to justify a true experiment.

In either case, quasi-experimental designs allow you to study the question by taking advantage of data that has previously been paid for or collected by others (often the government).

Quasi-experimental designs have various pros and cons compared to other types of studies.

  • Higher external validity than most true experiments, because they often involve real-world interventions instead of artificial laboratory settings.
  • Higher internal validity than other non-experimental types of research, because they allow you to better control for confounding variables than other types of studies do.
  • Lower internal validity than true experiments—without randomization, it can be difficult to verify that all confounding variables have been accounted for.
  • The use of retrospective data that has already been collected for other purposes can be inaccurate, incomplete or difficult to access.

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

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

Research bias

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

A quasi-experiment is a type of research design that attempts to establish a cause-and-effect relationship. The main difference with a true experiment is that the groups are not randomly assigned.

In experimental research, random assignment is a way of placing participants from your sample into different groups using randomization. With this method, every member of the sample has a known or equal chance of being placed in a control group or an experimental group.

Quasi-experimental design is most useful in situations where it would be unethical or impractical to run a true experiment .

Quasi-experiments have lower internal validity than true experiments, but they often have higher external validity  as they can use real-world interventions instead of artificial laboratory settings.

Cite this Scribbr article

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

Thomas, L. (2024, January 22). Quasi-Experimental Design | Definition, Types & Examples. Scribbr. Retrieved September 18, 2024, from https://www.scribbr.com/methodology/quasi-experimental-design/

Is this article helpful?

Lauren Thomas

Lauren Thomas

Other students also liked, guide to experimental design | overview, steps, & examples, random assignment in experiments | introduction & examples, control variables | what are they & why do they matter, "i thought ai proofreading was useless but..".

I've been using Scribbr for years now and I know it's a service that won't disappoint. It does a good job spotting mistakes”

  • Skip to main content
  • Skip to primary sidebar
  • Skip to footer
  • QuestionPro

survey software icon

  • Solutions Industries Gaming Automotive Sports and events Education Government Travel & Hospitality Financial Services Healthcare Cannabis Technology Use Case AskWhy Communities Audience Contactless surveys Mobile LivePolls Member Experience GDPR Positive People Science 360 Feedback Surveys
  • Resources Blog eBooks Survey Templates Case Studies Training Help center

quasi experimental definition in research

Home Market Research Research Tools and Apps

Quasi-experimental Research: What It Is, Types & Examples

quasi-experimental research is research that appears to be experimental but is not.

Much like an actual experiment, quasi-experimental research tries to demonstrate a cause-and-effect link between a dependent and an independent variable. A quasi-experiment, on the other hand, does not depend on random assignment, unlike an actual experiment. The subjects are sorted into groups based on non-random variables.

What is Quasi-Experimental Research?

“Resemblance” is the definition of “quasi.” Individuals are not randomly allocated to conditions or orders of conditions, even though the regression analysis is changed. As a result, quasi-experimental research is research that appears to be experimental but is not.

The directionality problem is avoided in quasi-experimental research since the regression analysis is altered before the multiple regression is assessed. However, because individuals are not randomized at random, there are likely to be additional disparities across conditions in quasi-experimental research.

As a result, in terms of internal consistency, quasi-experiments fall somewhere between correlational research and actual experiments.

The key component of a true experiment is randomly allocated groups. This means that each person has an equivalent chance of being assigned to the experimental group or the control group, depending on whether they are manipulated or not.

Simply put, a quasi-experiment is not a real experiment. A quasi-experiment does not feature randomly allocated groups since the main component of a real experiment is randomly assigned groups. Why is it so crucial to have randomly allocated groups, given that they constitute the only distinction between quasi-experimental and actual  experimental research ?

Let’s use an example to illustrate our point. Let’s assume we want to discover how new psychological therapy affects depressed patients. In a genuine trial, you’d split half of the psych ward into treatment groups, With half getting the new psychotherapy therapy and the other half receiving standard  depression treatment .

And the physicians compare the outcomes of this treatment to the results of standard treatments to see if this treatment is more effective. Doctors, on the other hand, are unlikely to agree with this genuine experiment since they believe it is unethical to treat one group while leaving another untreated.

A quasi-experimental study will be useful in this case. Instead of allocating these patients at random, you uncover pre-existing psychotherapist groups in the hospitals. Clearly, there’ll be counselors who are eager to undertake these trials as well as others who prefer to stick to the old ways.

These pre-existing groups can be used to compare the symptom development of individuals who received the novel therapy with those who received the normal course of treatment, even though the groups weren’t chosen at random.

If any substantial variations between them can be well explained, you may be very assured that any differences are attributable to the treatment but not to other extraneous variables.

As we mentioned before, quasi-experimental research entails manipulating an independent variable by randomly assigning people to conditions or sequences of conditions. Non-equivalent group designs, pretest-posttest designs, and regression discontinuity designs are only a few of the essential types.

What are quasi-experimental research designs?

Quasi-experimental research designs are a type of research design that is similar to experimental designs but doesn’t give full control over the independent variable(s) like true experimental designs do.

In a quasi-experimental design, the researcher changes or watches an independent variable, but the participants are not put into groups at random. Instead, people are put into groups based on things they already have in common, like their age, gender, or how many times they have seen a certain stimulus.

Because the assignments are not random, it is harder to draw conclusions about cause and effect than in a real experiment. However, quasi-experimental designs are still useful when randomization is not possible or ethical.

The true experimental design may be impossible to accomplish or just too expensive, especially for researchers with few resources. Quasi-experimental designs enable you to investigate an issue by utilizing data that has already been paid for or gathered by others (often the government). 

Because they allow better control for confounding variables than other forms of studies, they have higher external validity than most genuine experiments and higher  internal validity  (less than true experiments) than other non-experimental research.

Is quasi-experimental research quantitative or qualitative?

Quasi-experimental research is a quantitative research method. It involves numerical data collection and statistical analysis. Quasi-experimental research compares groups with different circumstances or treatments to find cause-and-effect links. 

It draws statistical conclusions from quantitative data. Qualitative data can enhance quasi-experimental research by revealing participants’ experiences and opinions, but quantitative data is the method’s foundation.

Quasi-experimental research types

There are many different sorts of quasi-experimental designs. Three of the most popular varieties are described below: Design of non-equivalent groups, Discontinuity in regression, and Natural experiments.

Design of Non-equivalent Groups

Example: design of non-equivalent groups, discontinuity in regression, example: discontinuity in regression, natural experiments, example: natural experiments.

However, because they couldn’t afford to pay everyone who qualified for the program, they had to use a random lottery to distribute slots.

Experts were able to investigate the program’s impact by utilizing enrolled people as a treatment group and those who were qualified but did not play the jackpot as an experimental group.

How QuestionPro helps in quasi-experimental research?

QuestionPro can be a useful tool in quasi-experimental research because it includes features that can assist you in designing and analyzing your research study. Here are some ways in which QuestionPro can help in quasi-experimental research:

Design surveys

Randomize participants, collect data over time, analyze data, collaborate with your team.

With QuestionPro, you have access to the most mature market research platform and tool that helps you collect and analyze the insights that matter the most. By leveraging InsightsHub, the unified hub for data management, you can ​​leverage the consolidated platform to organize, explore, search, and discover your  research data  in one organized data repository . 

Optimize Your quasi-experimental research with QuestionPro. Get started now!

LEARN MORE         FREE TRIAL

MORE LIKE THIS

Top 5 change management models to transform your organization.

Sep 20, 2024

customer reviews

Customer Reviews: How to Ask and Easy Ways to Get Them

Sep 19, 2024

Raw Data

Raw Data: What it is + How to Process It

Sep 18, 2024

quasi experimental definition in research

QuestionPro: Leading the Charge in Customer Journey Management and Voice of the Customer Platforms

Sep 17, 2024

Other categories

  • Academic Research
  • Artificial Intelligence
  • Assessments
  • Brand Awareness
  • Case Studies
  • Communities
  • Consumer Insights
  • Customer effort score
  • Customer Engagement
  • Customer Experience
  • Customer Loyalty
  • Customer Research
  • Customer Satisfaction
  • Employee Benefits
  • Employee Engagement
  • Employee Retention
  • Friday Five
  • General Data Protection Regulation
  • Insights Hub
  • Life@QuestionPro
  • Market Research
  • Mobile diaries
  • Mobile Surveys
  • New Features
  • Online Communities
  • Question Types
  • Questionnaire
  • QuestionPro Products
  • Release Notes
  • Research Tools and Apps
  • Revenue at Risk
  • Survey Templates
  • Training Tips
  • Tuesday CX Thoughts (TCXT)
  • Uncategorized
  • What’s Coming Up
  • Workforce Intelligence

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.

  • 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

Ethnographic Research

Ethnographic Research -Types, Methods and Guide

Mixed Research methods

Mixed Methods Research – Types & Analysis

Focus Groups in Qualitative Research

Focus Groups – Steps, Examples and Guide

Survey Research

Survey Research – Types, Methods, Examples

Transformative Design

Transformative Design – Methods, Types, Guide

Descriptive Research Design

Descriptive Research Design – Types, Methods and...

  • Skip to secondary menu
  • Skip to main content
  • Skip to primary sidebar

Statistics By Jim

Making statistics intuitive

Quasi Experimental Design Overview & Examples

By Jim Frost Leave a Comment

What is a Quasi Experimental Design?

A quasi experimental design is a method for identifying causal relationships that does not randomly assign participants to the experimental groups. Instead, researchers use a non-random process. For example, they might use an eligibility cutoff score or preexisting groups to determine who receives the treatment.

Image illustrating a quasi experimental design.

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 experiments. In these studies, researchers can manipulate an independent variable — that is, they change one factor to see what effect it has. However, unlike true experimental research, participants are not randomly assigned to different groups.

Learn more about Experimental Designs: Definition & Types .

When to Use Quasi-Experimental Design

Researchers typically use a quasi-experimental design because they can’t randomize due to practical or ethical concerns. For example:

  • Practical Constraints : A school interested in testing a new teaching method can only implement it in preexisting classes and cannot randomly assign students.
  • Ethical Concerns : A medical study might not be able to randomly assign participants to a treatment group for an experimental medication when they are already taking a proven drug.

Quasi-experimental designs also come in handy when researchers want to study the effects of naturally occurring events, like policy changes or environmental shifts, where they can’t control who is exposed to the treatment.

Quasi-experimental designs occupy a unique position in the spectrum of research methodologies, sitting between observational studies and true experiments. This middle ground offers a blend of both worlds, addressing some limitations of purely observational studies while navigating the constraints often accompanying true experiments.

A significant advantage of quasi-experimental research over purely observational studies and correlational research is that it addresses the issue of directionality, determining which variable is the cause and which is the effect. In quasi-experiments, an intervention typically occurs during the investigation, and the researchers record outcomes before and after it, increasing the confidence that it causes the observed changes.

However, it’s crucial to recognize its limitations as well. Controlling confounding variables is a larger concern for a quasi-experimental design than a true experiment because it lacks random assignment.

In sum, quasi-experimental designs offer a valuable research approach when random assignment is not feasible, providing a more structured and controlled framework than observational studies while acknowledging and attempting to address potential confounders.

Types of Quasi-Experimental Designs and Examples

Quasi-experimental studies use various methods, depending on the scenario.

Natural Experiments

This design uses naturally occurring events or changes to create the treatment and control groups. Researchers compare outcomes between those whom the event affected and those it did not affect. Analysts use statistical controls to account for confounders that the researchers must also measure.

Natural experiments are related to observational studies, but they allow for a clearer causality inference because the external event or policy change provides both a form of quasi-random group assignment and a definite start date for the intervention.

For example, in a natural experiment utilizing a quasi-experimental design, researchers study the impact of a significant economic policy change on small business growth. The policy is implemented in one state but not in neighboring states. This scenario creates an unplanned experimental setup, where the state with the new policy serves as the treatment group, and the neighboring states act as the control group.

Researchers are primarily interested in small business growth rates but need to record various confounders that can impact growth rates. Hence, they record state economic indicators, investment levels, and employment figures. By recording these metrics across the states, they can include them in the model as covariates and control them statistically. This method allows researchers to estimate differences in small business growth due to the policy itself, separate from the various confounders.

Nonequivalent Groups Design

This method involves matching existing groups that are similar but not identical. Researchers attempt to find groups that are as equivalent as possible, particularly for factors likely to affect the outcome.

For instance, researchers use a nonequivalent groups quasi-experimental design to evaluate the effectiveness of a new teaching method in improving students’ mathematics performance. A school district considering the teaching method is planning the study. Students are already divided into schools, preventing random assignment.

The researchers matched two schools with similar demographics, baseline academic performance, and resources. The school using the traditional methodology is the control, while the other uses the new approach. Researchers are evaluating differences in educational outcomes between the two methods.

They perform a pretest to identify differences between the schools that might affect the outcome and include them as covariates to control for confounding. They also record outcomes before and after the intervention to have a larger context for the changes they observe.

Regression Discontinuity

This process assigns subjects to a treatment or control group based on a predetermined cutoff point (e.g., a test score). The analysis primarily focuses on participants near the cutoff point, as they are likely similar except for the treatment received. By comparing participants just above and below the cutoff, the design controls for confounders that vary smoothly around the cutoff.

For example, in a regression discontinuity quasi-experimental design focusing on a new medical treatment for depression, researchers use depression scores as the cutoff point. Individuals with depression scores just above a certain threshold are assigned to receive the latest treatment, while those just below the threshold do not receive it. This method creates two closely matched groups: one that barely qualifies for treatment and one that barely misses out.

By comparing the mental health outcomes of these two groups over time, researchers can assess the effectiveness of the new treatment. The assumption is that the only significant difference between the groups is whether they received the treatment, thereby isolating its impact on depression outcomes.

Controlling Confounders in a Quasi-Experimental Design

Accounting for confounding variables is a challenging but essential task for a quasi-experimental design.

In a true experiment, the random assignment process equalizes confounders across the groups to nullify their overall effect. It’s the gold standard because it works on all confounders, known and unknown.

Unfortunately, the lack of random assignment can allow differences between the groups to exist before the intervention. These confounding factors might ultimately explain the results rather than the intervention.

Consequently, researchers must use other methods to equalize the groups roughly using matching and cutoff values or statistically adjust for preexisting differences they measure to reduce the impact of confounders.

A key strength of quasi-experiments is their frequent use of “pre-post testing.” This approach involves conducting initial tests before collecting data to check for preexisting differences between groups that could impact the study’s outcome. By identifying these variables early on and including them as covariates, researchers can more effectively control potential confounders in their statistical analysis.

Additionally, researchers frequently track outcomes before and after the intervention to better understand the context for changes they observe.

Statisticians consider these methods to be less effective than randomization. Hence, quasi-experiments fall somewhere in the middle when it comes to internal validity , or how well the study can identify causal relationships versus mere correlation . They’re more conclusive than correlational studies but not as solid as true experiments.

In conclusion, quasi-experimental designs offer researchers a versatile and practical approach when random assignment is not feasible. This methodology bridges the gap between controlled experiments and observational studies, providing a valuable tool for investigating cause-and-effect relationships in real-world settings. Researchers can address ethical and logistical constraints by understanding and leveraging the different types of quasi-experimental designs while still obtaining insightful and meaningful results.

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

Share this:

quasi experimental definition in research

Reader Interactions

Comments and questions cancel reply.

Quasi-Experimental Design: Definition, Types, Examples

Appinio Research · 19.12.2023 · 37min read

Quasi-Experimental Design Definition Types Examples

Ever wondered how researchers uncover cause-and-effect relationships in the real world, where controlled experiments are often elusive? Quasi-experimental design holds the key. In this guide, we'll unravel the intricacies of quasi-experimental design, shedding light on its definition, purpose, and applications across various domains. Whether you're a student, a professional, or simply curious about the methods behind meaningful research, join us as we delve into the world of quasi-experimental design, making complex concepts sound simple and embarking on a journey of knowledge and discovery.

What is Quasi-Experimental Design?

Quasi-experimental design is a research methodology used to study the effects of independent variables on dependent variables when full experimental control is not possible or ethical. It falls between controlled experiments, where variables are tightly controlled, and purely observational studies, where researchers have little control over variables. Quasi-experimental design mimics some aspects of experimental research but lacks randomization.

The primary purpose of quasi-experimental design is to investigate cause-and-effect relationships between variables in real-world settings. Researchers use this approach to answer research questions, test hypotheses, and explore the impact of interventions or treatments when they cannot employ traditional experimental methods. Quasi-experimental studies aim to maximize internal validity and make meaningful inferences while acknowledging practical constraints and ethical considerations.

Quasi-Experimental vs. Experimental Design

It's essential to understand the distinctions between Quasi-Experimental and Experimental Design to appreciate the unique characteristics of each approach:

  • Randomization:  In Experimental Design, random assignment of participants to groups is a defining feature. Quasi-experimental design, on the other hand, lacks randomization due to practical constraints or ethical considerations.
  • Control Groups :  Experimental Design typically includes control groups that are subjected to no treatment or a placebo. The quasi-experimental design may have comparison groups but lacks the same level of control.
  • Manipulation of IV:  Experimental Design involves the intentional manipulation of the independent variable. Quasi-experimental design often deals with naturally occurring independent variables.
  • Causal Inference:  Experimental Design allows for stronger causal inferences due to randomization and control. Quasi-experimental design permits causal inferences but with some limitations.

When to Use Quasi-Experimental Design?

A quasi-experimental design is particularly valuable in several situations:

  • Ethical Constraints:  When manipulating the independent variable is ethically unacceptable or impractical, quasi-experimental design offers an alternative to studying naturally occurring variables.
  • Real-World Settings:  When researchers want to study phenomena in real-world contexts, quasi-experimental design allows them to do so without artificial laboratory settings.
  • Limited Resources:  In cases where resources are limited and conducting a controlled experiment is cost-prohibitive, quasi-experimental design can provide valuable insights.
  • Policy and Program Evaluation:  Quasi-experimental design is commonly used in evaluating the effectiveness of policies, interventions, or programs that cannot be randomly assigned to participants.

Importance of Quasi-Experimental Design in Research

Quasi-experimental design plays a vital role in research for several reasons:

  • Addressing Real-World Complexities:  It allows researchers to tackle complex real-world issues where controlled experiments are not feasible. This bridges the gap between controlled experiments and purely observational studies.
  • Ethical Research:  It provides an honest approach when manipulating variables or assigning treatments could harm participants or violate ethical standards.
  • Policy and Practice Implications:  Quasi-experimental studies generate findings with direct applications in policy-making and practical solutions in fields such as education, healthcare, and social sciences.
  • Enhanced External Validity:  Findings from Quasi-Experimental research often have high external validity, making them more applicable to broader populations and contexts.

By embracing the challenges and opportunities of quasi-experimental design, researchers can contribute valuable insights to their respective fields and drive positive changes in the real world.

Key Concepts in Quasi-Experimental Design

In quasi-experimental design, it's essential to grasp the fundamental concepts underpinning this research methodology. Let's explore these key concepts in detail.

Independent Variable

The independent variable (IV) is the factor you aim to study or manipulate in your research. Unlike controlled experiments, where you can directly manipulate the IV, quasi-experimental design often deals with naturally occurring variables. For example, if you're investigating the impact of a new teaching method on student performance, the teaching method is your independent variable.

Dependent Variable

The dependent variable (DV) is the outcome or response you measure to assess the effects of changes in the independent variable. Continuing with the teaching method example, the dependent variable would be the students' academic performance, typically measured using test scores, grades, or other relevant metrics.

Control Groups vs. Comparison Groups

While quasi-experimental design lacks the luxury of randomly assigning participants to control and experimental groups, you can still establish comparison groups to make meaningful inferences. Control groups consist of individuals who do not receive the treatment, while comparison groups are exposed to different levels or variations of the treatment. These groups help researchers gauge the effect of the independent variable.

Pre-Test and Post-Test Measures

In quasi-experimental design, it's common practice to collect data both before and after implementing the independent variable. The initial data (pre-test) serves as a baseline, allowing you to measure changes over time (post-test). This approach helps assess the impact of the independent variable more accurately. For instance, if you're studying the effectiveness of a new drug, you'd measure patients' health before administering the drug (pre-test) and afterward (post-test).

Threats to Internal Validity

Internal validity is crucial for establishing a cause-and-effect relationship between the independent and dependent variables. However, in a quasi-experimental design, several threats can compromise internal validity. These threats include:

  • Selection Bias :  When non-randomized groups differ systematically in ways that affect the study's outcome.
  • History Effects:  External events or changes over time that influence the results.
  • Maturation Effects:  Natural changes or developments that occur within participants during the study.
  • Regression to the Mean:  The tendency for extreme scores on a variable to move closer to the mean upon retesting.
  • Attrition and Mortality:  The loss of participants over time, potentially skewing the results.
  • Testing Effects:  The mere act of testing or assessing participants can impact their subsequent performance.

Understanding these threats is essential for designing and conducting Quasi-Experimental studies that yield valid and reliable results.

Randomization and Non-Randomization

In traditional experimental designs, randomization is a powerful tool for ensuring that groups are equivalent at the outset of a study. However, quasi-experimental design often involves non-randomization due to the nature of the research. This means that participants are not randomly assigned to treatment and control groups. Instead, researchers must employ various techniques to minimize biases and ensure that the groups are as similar as possible.

For example, if you are conducting a study on the effects of a new teaching method in a real classroom setting, you cannot randomly assign students to the treatment and control groups. Instead, you might use statistical methods to match students based on relevant characteristics such as prior academic performance or socioeconomic status. This matching process helps control for potential confounding variables, increasing the validity of your study.

Types of Quasi-Experimental Designs

In quasi-experimental design, researchers employ various approaches to investigate causal relationships and study the effects of independent variables when complete experimental control is challenging. Let's explore these types of quasi-experimental designs.

One-Group Posttest-Only Design

The One-Group Posttest-Only Design is one of the simplest forms of quasi-experimental design. In this design, a single group is exposed to the independent variable, and data is collected only after the intervention has taken place. Unlike controlled experiments, there is no comparison group. This design is useful when researchers cannot administer a pre-test or when it is logistically difficult to do so.

Example : Suppose you want to assess the effectiveness of a new time management seminar. You offer the seminar to a group of employees and measure their productivity levels immediately afterward to determine if there's an observable impact.

One-Group Pretest-Posttest Design

Similar to the One-Group Posttest-Only Design, this approach includes a pre-test measure in addition to the post-test. Researchers collect data both before and after the intervention. By comparing the pre-test and post-test results within the same group, you can gain a better understanding of the changes that occur due to the independent variable.

Example : If you're studying the impact of a stress management program on participants' stress levels, you would measure their stress levels before the program (pre-test) and after completing the program (post-test) to assess any changes.

Non-Equivalent Groups Design

The Non-Equivalent Groups Design involves multiple groups, but they are not randomly assigned. Instead, researchers must carefully match or control for relevant variables to minimize biases. This design is particularly useful when random assignment is not possible or ethical.

Example : Imagine you're examining the effectiveness of two teaching methods in two different schools. You can't randomly assign students to the schools, but you can carefully match them based on factors like age, prior academic performance, and socioeconomic status to create equivalent groups.

Time Series Design

Time Series Design is an approach where data is collected at multiple time points before and after the intervention. This design allows researchers to analyze trends and patterns over time, providing valuable insights into the sustained effects of the independent variable.

Example : If you're studying the impact of a new marketing campaign on product sales, you would collect sales data at regular intervals (e.g., monthly) before and after the campaign's launch to observe any long-term trends.

Regression Discontinuity Design

Regression Discontinuity Design is employed when participants are assigned to different groups based on a specific cutoff score or threshold. This design is often used in educational and policy research to assess the effects of interventions near a cutoff point.

Example : Suppose you're evaluating the impact of a scholarship program on students' academic performance. Students who score just above or below a certain GPA threshold are assigned differently to the program. This design helps assess the program's effectiveness at the cutoff point.

Propensity Score Matching

Propensity Score Matching is a technique used to create comparable treatment and control groups in non-randomized studies. Researchers calculate propensity scores based on participants' characteristics and match individuals in the treatment group to those in the control group with similar scores.

Example : If you're studying the effects of a new medication on patient outcomes, you would use propensity scores to match patients who received the medication with those who did not but have similar health profiles.

Interrupted Time Series Design

The Interrupted Time Series Design involves collecting data at multiple time points before and after the introduction of an intervention. However, in this design, the intervention occurs at a specific point in time, allowing researchers to assess its immediate impact.

Example : Let's say you're analyzing the effects of a new traffic management system on traffic accidents. You collect accident data before and after the system's implementation to observe any abrupt changes right after its introduction.

Each of these quasi-experimental designs offers unique advantages and is best suited to specific research questions and scenarios. Choosing the right design is crucial for conducting robust and informative studies.

Advantages and Disadvantages of Quasi-Experimental Design

Quasi-experimental design offers a valuable research approach, but like any methodology, it comes with its own set of advantages and disadvantages. Let's explore these in detail.

Quasi-Experimental Design Advantages

Quasi-experimental design presents several advantages that make it a valuable tool in research:

  • Real-World Applicability:  Quasi-experimental studies often take place in real-world settings, making the findings more applicable to practical situations. Researchers can examine the effects of interventions or variables in the context where they naturally occur.
  • Ethical Considerations:  In situations where manipulating the independent variable in a controlled experiment would be unethical, quasi-experimental design provides an ethical alternative. For example, it would be unethical to assign participants to smoke for a study on the health effects of smoking, but you can study naturally occurring groups of smokers and non-smokers.
  • Cost-Efficiency:  Conducting Quasi-Experimental research is often more cost-effective than conducting controlled experiments. The absence of controlled environments and extensive manipulations can save both time and resources.

These advantages make quasi-experimental design an attractive choice for researchers facing practical or ethical constraints in their studies.

Quasi-Experimental Design Disadvantages

However, quasi-experimental design also comes with its share of challenges and disadvantages:

  • Limited Control:  Unlike controlled experiments, where researchers have full control over variables, quasi-experimental design lacks the same level of control. This limited control can result in confounding variables that make it difficult to establish causality.
  • Threats to Internal Validity:  Various threats to internal validity, such as selection bias, history effects, and maturation effects, can compromise the accuracy of causal inferences. Researchers must carefully address these threats to ensure the validity of their findings.
  • Causality Inference Challenges:  Establishing causality can be challenging in quasi-experimental design due to the absence of randomization and control. While you can make strong arguments for causality, it may not be as conclusive as in controlled experiments.
  • Potential Confounding Variables:  In a quasi-experimental design, it's often challenging to control for all possible confounding variables that may affect the dependent variable. This can lead to uncertainty in attributing changes solely to the independent variable.

Despite these disadvantages, quasi-experimental design remains a valuable research tool when used judiciously and with a keen awareness of its limitations. Researchers should carefully consider their research questions and the practical constraints they face before choosing this approach.

How to Conduct a Quasi-Experimental Study?

Conducting a Quasi-Experimental study requires careful planning and execution to ensure the validity of your research. Let's dive into the essential steps you need to follow when conducting such a study.

1. Define Research Questions and Objectives

The first step in any research endeavor is clearly defining your research questions and objectives. This involves identifying the independent variable (IV) and the dependent variable (DV) you want to study. What is the specific relationship you want to explore, and what do you aim to achieve with your research?

  • Specify Your Research Questions :  Start by formulating precise research questions that your study aims to answer. These questions should be clear, focused, and relevant to your field of study.
  • Identify the Independent Variable:  Define the variable you intend to manipulate or study in your research. Understand its significance in your study's context.
  • Determine the Dependent Variable:  Identify the outcome or response variable that will be affected by changes in the independent variable.
  • Establish Hypotheses (If Applicable):  If you have specific hypotheses about the relationship between the IV and DV, state them clearly. Hypotheses provide a framework for testing your research questions.

2. Select the Appropriate Quasi-Experimental Design

Choosing the right quasi-experimental design is crucial for achieving your research objectives. Select a design that aligns with your research questions and the available data. Consider factors such as the feasibility of implementing the design and the ethical considerations involved.

  • Evaluate Your Research Goals:  Assess your research questions and objectives to determine which type of quasi-experimental design is most suitable. Each design has its strengths and limitations, so choose one that aligns with your goals.
  • Consider Ethical Constraints:  Take into account any ethical concerns related to your research. Depending on your study's context, some designs may be more ethically sound than others.
  • Assess Data Availability:  Ensure you have access to the necessary data for your chosen design. Some designs may require extensive historical data, while others may rely on data collected during the study.

3. Identify and Recruit Participants

Selecting the right participants is a critical aspect of Quasi-Experimental research. The participants should represent the population you want to make inferences about, and you must address ethical considerations, including informed consent.

  • Define Your Target Population:  Determine the population that your study aims to generalize to. Your sample should be representative of this population.
  • Recruitment Process:  Develop a plan for recruiting participants. Depending on your design, you may need to reach out to specific groups or institutions.
  • Informed Consent:  Ensure that you obtain informed consent from participants. Clearly explain the nature of the study, potential risks, and their rights as participants.

4. Collect Data

Data collection is a crucial step in Quasi-Experimental research. You must adhere to a consistent and systematic process to gather relevant information before and after the intervention or treatment.

  • Pre-Test Measures:  If applicable, collect data before introducing the independent variable. Ensure that the pre-test measures are standardized and reliable.
  • Post-Test Measures:  After the intervention, collect post-test data using the same measures as the pre-test. This allows you to assess changes over time.
  • Maintain Data Consistency:  Ensure that data collection procedures are consistent across all participants and time points to minimize biases.

5. Analyze Data

Once you've collected your data, it's time to analyze it using appropriate statistical techniques . The choice of analysis depends on your research questions and the type of data you've gathered.

  • Statistical Analysis :  Use statistical software to analyze your data. Common techniques include t-tests , analysis of variance (ANOVA) , regression analysis , and more, depending on the design and variables.
  • Control for Confounding Variables:  Be aware of potential confounding variables and include them in your analysis as covariates to ensure accurate results.

Chi-Square Calculator :

t-Test Calculator :

6. Interpret Results

With the analysis complete, you can interpret the results to draw meaningful conclusions about the relationship between the independent and dependent variables.

  • Examine Effect Sizes:  Assess the magnitude of the observed effects to determine their practical significance.
  • Consider Significance Levels:  Determine whether the observed results are statistically significant . Understand the p-values and their implications.
  • Compare Findings to Hypotheses:  Evaluate whether your findings support or reject your hypotheses and research questions.

7. Draw Conclusions

Based on your analysis and interpretation of the results, draw conclusions about the research questions and objectives you set out to address.

  • Causal Inferences:  Discuss the extent to which your study allows for causal inferences. Be transparent about the limitations and potential alternative explanations for your findings.
  • Implications and Applications:  Consider the practical implications of your research. How do your findings contribute to existing knowledge, and how can they be applied in real-world contexts?
  • Future Research:  Identify areas for future research and potential improvements in study design. Highlight any limitations or constraints that may have affected your study's outcomes.

By following these steps meticulously, you can conduct a rigorous and informative Quasi-Experimental study that advances knowledge in your field of research.

Quasi-Experimental Design Examples

Quasi-experimental design finds applications in a wide range of research domains, including business-related and market research scenarios. Below, we delve into some detailed examples of how this research methodology is employed in practice:

Example 1: Assessing the Impact of a New Marketing Strategy

Suppose a company wants to evaluate the effectiveness of a new marketing strategy aimed at boosting sales. Conducting a controlled experiment may not be feasible due to the company's existing customer base and the challenge of randomly assigning customers to different marketing approaches. In this scenario, a quasi-experimental design can be employed.

  • Independent Variable:  The new marketing strategy.
  • Dependent Variable:  Sales revenue.
  • Design:  The company could implement the new strategy for one group of customers while maintaining the existing strategy for another group. Both groups are selected based on similar demographics and purchase history , reducing selection bias. Pre-implementation data (sales records) can serve as the baseline, and post-implementation data can be collected to assess the strategy's impact.

Example 2: Evaluating the Effectiveness of Employee Training Programs

In the context of human resources and employee development, organizations often seek to evaluate the impact of training programs. A randomized controlled trial (RCT) with random assignment may not be practical or ethical, as some employees may need specific training more than others. Instead, a quasi-experimental design can be employed.

  • Independent Variable:  Employee training programs.
  • Dependent Variable:  Employee performance metrics, such as productivity or quality of work.
  • Design:  The organization can offer training programs to employees who express interest or demonstrate specific needs, creating a self-selected treatment group. A comparable control group can consist of employees with similar job roles and qualifications who did not receive the training. Pre-training performance metrics can serve as the baseline, and post-training data can be collected to assess the impact of the training programs.

Example 3: Analyzing the Effects of a Tax Policy Change

In economics and public policy, researchers often examine the effects of tax policy changes on economic behavior. Conducting a controlled experiment in such cases is practically impossible. Therefore, a quasi-experimental design is commonly employed.

  • Independent Variable:  Tax policy changes (e.g., tax rate adjustments).
  • Dependent Variable:  Economic indicators, such as consumer spending or business investments.
  • Design:  Researchers can analyze data from different regions or jurisdictions where tax policy changes have been implemented. One region could represent the treatment group (with tax policy changes), while a similar region with no tax policy changes serves as the control group. By comparing economic data before and after the policy change in both groups, researchers can assess the impact of the tax policy changes.

These examples illustrate how quasi-experimental design can be applied in various research contexts, providing valuable insights into the effects of independent variables in real-world scenarios where controlled experiments are not feasible or ethical. By carefully selecting comparison groups and controlling for potential biases, researchers can draw meaningful conclusions and inform decision-making processes.

How to Publish Quasi-Experimental Research?

Publishing your Quasi-Experimental research findings is a crucial step in contributing to the academic community's knowledge. We'll explore the essential aspects of reporting and publishing your Quasi-Experimental research effectively.

Structuring Your Research Paper

When preparing your research paper, it's essential to adhere to a well-structured format to ensure clarity and comprehensibility. Here are key elements to include:

Title and Abstract

  • Title:  Craft a concise and informative title that reflects the essence of your study. It should capture the main research question or hypothesis.
  • Abstract:  Summarize your research in a structured abstract, including the purpose, methods, results, and conclusions. Ensure it provides a clear overview of your study.

Introduction

  • Background and Rationale:  Provide context for your study by discussing the research gap or problem your study addresses. Explain why your research is relevant and essential.
  • Research Questions or Hypotheses:  Clearly state your research questions or hypotheses and their significance.

Literature Review

  • Review of Related Work:  Discuss relevant literature that supports your research. Highlight studies with similar methodologies or findings and explain how your research fits within this context.
  • Participants:  Describe your study's participants, including their characteristics and how you recruited them.
  • Quasi-Experimental Design:  Explain your chosen design in detail, including the independent and dependent variables, procedures, and any control measures taken.
  • Data Collection:  Detail the data collection methods , instruments used, and any pre-test or post-test measures.
  • Data Analysis:  Describe the statistical techniques employed, including any control for confounding variables.
  • Presentation of Findings:  Present your results clearly, using tables, graphs, and descriptive statistics where appropriate. Include p-values and effect sizes, if applicable.
  • Interpretation of Results:  Discuss the implications of your findings and how they relate to your research questions or hypotheses.
  • Interpretation and Implications:  Analyze your results in the context of existing literature and theories. Discuss the practical implications of your findings.
  • Limitations:  Address the limitations of your study, including potential biases or threats to internal validity.
  • Future Research:  Suggest areas for future research and how your study contributes to the field.

Ethical Considerations in Reporting

Ethical reporting is paramount in Quasi-Experimental research. Ensure that you adhere to ethical standards, including:

  • Informed Consent:  Clearly state that informed consent was obtained from all participants, and describe the informed consent process.
  • Protection of Participants:  Explain how you protected the rights and well-being of your participants throughout the study.
  • Confidentiality:  Detail how you maintained privacy and anonymity, especially when presenting individual data.
  • Disclosure of Conflicts of Interest:  Declare any potential conflicts of interest that could influence the interpretation of your findings.

Common Pitfalls to Avoid

When reporting your Quasi-Experimental research, watch out for common pitfalls that can diminish the quality and impact of your work:

  • Overgeneralization:  Be cautious not to overgeneralize your findings. Clearly state the limits of your study and the populations to which your results can be applied.
  • Misinterpretation of Causality:  Clearly articulate the limitations in inferring causality in Quasi-Experimental research. Avoid making strong causal claims unless supported by solid evidence.
  • Ignoring Ethical Concerns:  Ethical considerations are paramount. Failing to report on informed consent, ethical oversight, and participant protection can undermine the credibility of your study.

Guidelines for Transparent Reporting

To enhance the transparency and reproducibility of your Quasi-Experimental research, consider adhering to established reporting guidelines, such as:

  • CONSORT Statement:  If your study involves interventions or treatments, follow the CONSORT guidelines for transparent reporting of randomized controlled trials.
  • STROBE Statement:  For observational studies, the STROBE statement provides guidance on reporting essential elements.
  • PRISMA Statement:  If your research involves systematic reviews or meta-analyses, adhere to the PRISMA guidelines.
  • Transparent Reporting of Evaluations with Non-Randomized Designs (TREND):  TREND guidelines offer specific recommendations for transparently reporting non-randomized designs, including Quasi-Experimental research.

By following these reporting guidelines and maintaining the highest ethical standards, you can contribute to the advancement of knowledge in your field and ensure the credibility and impact of your Quasi-Experimental research findings.

Quasi-Experimental Design Challenges

Conducting a Quasi-Experimental study can be fraught with challenges that may impact the validity and reliability of your findings. We'll take a look at some common challenges and provide strategies on how you can address them effectively.

Selection Bias

Challenge:  Selection bias occurs when non-randomized groups differ systematically in ways that affect the study's outcome. This bias can undermine the validity of your research, as it implies that the groups are not equivalent at the outset of the study.

Addressing Selection Bias:

  • Matching:  Employ matching techniques to create comparable treatment and control groups. Match participants based on relevant characteristics, such as age, gender, or prior performance, to balance the groups.
  • Statistical Controls:  Use statistical controls to account for differences between groups. Include covariates in your analysis to adjust for potential biases.
  • Sensitivity Analysis:  Conduct sensitivity analyses to assess how vulnerable your results are to selection bias. Explore different scenarios to understand the impact of potential bias on your conclusions.

History Effects

Challenge:  History effects refer to external events or changes over time that influence the study's results. These external factors can confound your research by introducing variables you did not account for.

Addressing History Effects:

  • Collect Historical Data:  Gather extensive historical data to understand trends and patterns that might affect your study. By having a comprehensive historical context, you can better identify and account for historical effects.
  • Control Groups:  Include control groups whenever possible. By comparing the treatment group's results to those of a control group, you can account for external influences that affect both groups equally.
  • Time Series Analysis :  If applicable, use time series analysis to detect and account for temporal trends. This method helps differentiate between the effects of the independent variable and external events.

Maturation Effects

Challenge:  Maturation effects occur when participants naturally change or develop throughout the study, independent of the intervention. These changes can confound your results, making it challenging to attribute observed effects solely to the independent variable.

Addressing Maturation Effects:

  • Randomization:  If possible, use randomization to distribute maturation effects evenly across treatment and control groups. Random assignment minimizes the impact of maturation as a confounding variable.
  • Matched Pairs:  If randomization is not feasible, employ matched pairs or statistical controls to ensure that both groups experience similar maturation effects.
  • Shorter Time Frames:  Limit the duration of your study to reduce the likelihood of significant maturation effects. Shorter studies are less susceptible to long-term maturation.

Regression to the Mean

Challenge:  Regression to the mean is the tendency for extreme scores on a variable to move closer to the mean upon retesting. This can create the illusion of an intervention's effectiveness when, in reality, it's a natural statistical phenomenon.

Addressing Regression to the Mean:

  • Use Control Groups:  Include control groups in your study to provide a baseline for comparison. This helps differentiate genuine intervention effects from regression to the mean.
  • Multiple Data Points:  Collect numerous data points to identify patterns and trends. If extreme scores regress to the mean in subsequent measurements, it may be indicative of regression to the mean rather than a true intervention effect.
  • Statistical Analysis:  Employ statistical techniques that account for regression to the mean when analyzing your data. Techniques like analysis of covariance (ANCOVA) can help control for baseline differences.

Attrition and Mortality

Challenge:  Attrition refers to the loss of participants over the course of your study, while mortality is the permanent loss of participants. High attrition rates can introduce biases and affect the representativeness of your sample.

Addressing Attrition and Mortality:

  • Careful Participant Selection:  Select participants who are likely to remain engaged throughout the study. Consider factors that may lead to attrition, such as participant motivation and commitment.
  • Incentives:  Provide incentives or compensation to participants to encourage their continued participation.
  • Follow-Up Strategies:  Implement effective follow-up strategies to reduce attrition. Regular communication and reminders can help keep participants engaged.
  • Sensitivity Analysis:  Conduct sensitivity analyses to assess the impact of attrition and mortality on your results. Compare the characteristics of participants who dropped out with those who completed the study.

Testing Effects

Challenge:  Testing effects occur when the mere act of testing or assessing participants affects their subsequent performance. This phenomenon can lead to changes in the dependent variable that are unrelated to the independent variable.

Addressing Testing Effects:

  • Counterbalance Testing:  If possible, counterbalance the order of tests or assessments between treatment and control groups. This helps distribute the testing effects evenly across groups.
  • Control Groups:  Include control groups subjected to the same testing or assessment procedures as the treatment group. By comparing the two groups, you can determine whether testing effects have influenced the results.
  • Minimize Testing Frequency:  Limit the frequency of testing or assessments to reduce the likelihood of testing effects. Conducting fewer assessments can mitigate the impact of repeated testing on participants.

By proactively addressing these common challenges, you can enhance the validity and reliability of your Quasi-Experimental study, making your findings more robust and trustworthy.

Conclusion for Quasi-Expermental Design

Quasi-experimental design is a powerful tool that helps researchers investigate cause-and-effect relationships in real-world situations where strict control is not always possible. By understanding the key concepts, types of designs, and how to address challenges, you can conduct robust research and contribute valuable insights to your field. Remember, quasi-experimental design bridges the gap between controlled experiments and purely observational studies, making it an essential approach in various fields, from business and market research to public policy and beyond. So, whether you're a researcher, student, or decision-maker, the knowledge of quasi-experimental design empowers you to make informed choices and drive positive changes in the world.

How to Supercharge Quasi-Experimental Design with Real-Time Insights?

Introducing Appinio , the real-time market research platform that transforms the world of quasi-experimental design. Imagine having the power to conduct your own market research in minutes, obtaining actionable insights that fuel your data-driven decisions. Appinio takes care of the research and tech complexities, freeing you to focus on what truly matters for your business.

Here's why Appinio stands out:

  • Lightning-Fast Insights:  From formulating questions to uncovering insights, Appinio delivers results in minutes, ensuring you get the answers you need when you need them.
  • No Research Degree Required:  Our intuitive platform is designed for everyone, eliminating the need for a PhD in research. Anyone can dive in and start harnessing the power of real-time consumer insights.
  • Global Reach, Local Expertise:  With access to over 90 countries and the ability to define precise target groups based on 1200+ characteristics, you can conduct Quasi-Experimental research on a global scale while maintaining a local touch.

Register now EN

Get free access to the platform!

Join the loop 💌

Be the first to hear about new updates, product news, and data insights. We'll send it all straight to your inbox.

Get the latest market research news straight to your inbox! 💌

Wait, there's more

Track Your Customer Retention & Brand Metrics for Post-Holiday Success

19.09.2024 | 8min read

Track Your Customer Retention & Brand Metrics for Post-Holiday Success

Creative Checkup – Optimize Advertising Slogans & Creatives for maximum ROI

16.09.2024 | 9min read

Creative Checkup – Optimize Advertising Slogans & Creatives for ROI

Get your brand Holiday Ready: 4 Essential Steps to Smash your Q4

03.09.2024 | 6min read

Get your brand Holiday Ready: 4 Essential Steps to Smash your Q4

Our systems are now restored following recent technical disruption, and we’re working hard to catch up on publishing. We apologise for the inconvenience caused. Find out more: https://www.cambridge.org/universitypress/about-us/news-and-blogs/cambridge-university-press-publishing-update-following-technical-disruption

We use cookies to distinguish you from other users and to provide you with a better experience on our websites. Close this message to accept cookies or find out how to manage your cookie settings .

Login Alert

  • > The Cambridge Handbook of Research Methods and Statistics for the Social and Behavioral Sciences
  • > Quasi-Experimental Research

quasi experimental definition in research

Book contents

  • The Cambridge Handbook of Research Methods and Statistics for the Social and Behavioral Sciences
  • Cambridge Handbooks in Psychology
  • Copyright page
  • Contributors
  • Part I From Idea to Reality: The Basics of Research
  • Part II The Building Blocks of a Study
  • Part III Data Collection
  • 13 Cross-Sectional Studies
  • 14 Quasi-Experimental Research
  • 15 Non-equivalent Control Group Pretest–Posttest Design in Social and Behavioral Research
  • 16 Experimental Methods
  • 17 Longitudinal Research: A World to Explore
  • 18 Online Research Methods
  • 19 Archival Data
  • 20 Qualitative Research Design
  • Part IV Statistical Approaches
  • Part V Tips for a Successful Research Career

14 - Quasi-Experimental Research

from Part III - Data Collection

Published online by Cambridge University Press:  25 May 2023

In this chapter, we discuss the logic and practice of quasi-experimentation. Specifically, we describe four quasi-experimental designs – one-group pretest–posttest designs, non-equivalent group designs, regression discontinuity designs, and interrupted time-series designs – and their statistical analyses in detail. Both simple quasi-experimental designs and embellishments of these simple designs are presented. Potential threats to internal validity are illustrated along with means of addressing their potentially biasing effects so that these effects can be minimized. In contrast to quasi-experiments, randomized experiments are often thought to be the gold standard when estimating the effects of treatment interventions. However, circumstances frequently arise where quasi-experiments can usefully supplement randomized experiments or when quasi-experiments can fruitfully be used in place of randomized experiments. Researchers need to appreciate the relative strengths and weaknesses of the various quasi-experiments so they can choose among pre-specified designs or craft their own unique quasi-experiments.

Access options

Save book to kindle.

To save this book to your Kindle, first ensure [email protected] is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle .

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service .

  • Quasi-Experimental Research
  • By Charles S. Reichardt , Daniel Storage , Damon Abraham
  • Edited by Austin Lee Nichols , Central European University, Vienna , John Edlund , Rochester Institute of Technology, New York
  • Book: The Cambridge Handbook of Research Methods and Statistics for the Social and Behavioral Sciences
  • Online publication: 25 May 2023
  • Chapter DOI: https://doi.org/10.1017/9781009010054.015

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox .

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive .

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: Sep 11, 2024 11:05 AM
  • URL: https://instr.iastate.libguides.com/researchmethods
  • Quasi-Experimental Research: Types, Examples & Application

Moradeke Owa

Let’s say you want to study the effects of a new drug on lowering blood pressure. You could randomly assign half of the participants to receive the drug and the other half to receive a placebo. However, this isn’t fair to the patients in the placebo group. 

So, instead of experimenting to find out the effect of the new drug, you use a quasi-experiment to evaluate the drug. The purpose of quasi-experimental research is to establish a causal relationship between an independent variable and a dependent variable.

This guide will discuss the different types of quasi-experimental research, their practical applications, and the best practices for conducting successful quasi-experimental research.

Understanding Quasi-Experimental Research

quasi experimental definition in research

Quasi-experimental research is a way of finding out if there’s a cause-and-effect relationship between variables when true experiments are not possible because of practical or ethical constraints.

For example, you want to know if a new medicine is effective for migraines. Instead of giving the medication to some people and not to others, you use quasi-experimental research to compare who takes the medication and people who haven’t.

What’s The Difference Between True Experiments and Quasi-Experiments

Quasi-experimental research doesn’t always have the same level of internal validity as real experiments. Pre-selected samples can be biased and not represent the real population.

In a true experiment, the participants are assigned randomly to the experimental or control groups. This ensures that groups are as homogeneous as possible, except for the treatment they receive.

Quasi-experiments don’t randomly assign participants to groups, so the differences within the groups could affect the groups.

Read – Experimental Research Designs: Types, Examples & Methods  

  • Types of Quasi-Experimental Designs

Pretest-Posttest Design

This design captures changes by measuring participants before and after an intervention. The pretest measures the dependent variable before the intervention, while the posttest measures it after the intervention.

The difference between the two measurements is the change that occurred due to the intervention. 

However, it is important to be aware of the potential threats to the internal validity of the pretest-postest design. One threat is selection bias , which happens when the group is not homogenous. 

Another is maturation, which occurs when participants change naturally over time. You can mitigate these threats using a control group, randomization, and blinding techniques.

quasi experimental definition in research

Posttest-Only Design with Nonequivalent Groups

The posttest-only design with nonequivalent groups is similar to the pretest-posttest design, but it does not include a pretest. You see the effect of the intervention by comparing groups to see if there is a difference in their scores.

The difference in scores determines the impact of the intervention. This design is less powerful than the pretest-posttest design because it does not account for potential differences between the groups at baseline. 

However, the posttest design is still a valuable tool for research, especially when you can’t collect pretest data. You can mitigate its limitations by using matching participants on important factors, statistical analysis, and sensitivity analysis

  • Regression Discontinuity Design

The regression discontinuity design uses naturally occurring cutoff points to assign participants to treatment and control groups. It’s widely adopted in education and social policy research. 

For example, a talent recruiter might use the cutoff score on a standardized test to move candidates to the next phase of the application.

Interrupted Time Series Design

The Interrupted Time Series design is a type of study that uses time series data to figure out how interventions or events affect the population. In this study, you measure the dependent variable multiple times over time, before and after the intervention.

The interrupted time series design is most commonly used to study the impact of policies or programs. For example, you are studying the impact of a new law on traffic accidents.

You could collect data on the number of traffic accidents each month for a year before the law was passed and a year after the law was passed. If the number of traffic accidents decreases after the law was passed, then you could conclude that the law had a positive impact on traffic safety.

Rigor in Quasi-Experimental Research

Matching techniques.

Matching techniques are a way to create balanced treatment and control groups in quasi-experimental research. This is done by matching participants on important characteristics, such as age, gender, or socioeconomic status.

Propensity score matching is one of the most popular matching methods. It works by using a statistical model to figure out how likely it is that each person in the study would have been in the treatment group if they were selected randomly. Then, people are randomly assigned according to their propensity scores, making sure that the treatment group and the control group are as close to the same as possible.

  • Creates balanced treatment and control groups, which reduces bias in the results.
  • Versatile- you can use them in various research settings, especially where randomization is not possible.
  • Relatively easy to implement.
  • Computationally complex.
  • Sensitive to the choice of the matching algorithm.
  • Does not perfectly balance the treatment and control groups on all relevant characteristics.

Instrumental Variables

quasi experimental definition in research

An instrumental variable (IV) in quasi-experimental research is a variable that’s related to the independent variable, but not to the error term. It’s a variable that can be used to measure how the independent variable affects the dependent variable.

Let’s say you want to investigate how a new drug reduces the risk of heart attack. You can use the number of days a person has taken aspirin as the instrumental variable.

Aspirin is associated with an independent variable (new drug), however, it is not associated with a dependent variable (risk of a heart attack). This is because people who take aspirin are more likely to take other medications, such as statins, which also lower the risk of heart attack.

  • Addresses endogeneity issues.
  • Versatile-you can use it in different research settings
  • It is relatively hard to find an instrumental variable that meets all of the criteria.
  • May not be perfectly correlated with the independent variable.
  • Tends to be affected by the dependent variable.

Difference-in-Differences Analysis

quasi experimental definition in research

Difference-in-differences analysis is a statistical technique that can be used to compare changes in treatment and control groups over time. It is typically used in quasi-experimental research to estimate the causal effect of an intervention.

You have to first define two groups when using the difference-in-differences analysis: the treatment group and the control group. A treatment group is a group that receives an intervention, while a control group doesn’t receive an intervention.

Next, collect data on the dependent variable for both groups before and after the intervention. The difference-in-differences estimate is then calculated by comparing the change in the dependent variable for the treatment group to the change in the dependent variable for the control group.

For example, in a study by David Card and Alan Krueger , they compared the effect of increasing the minimum wage in a particular region to the employment rate. They found that the minimum wage increase in New Jersey did not lead to job losses.

  • Addresses selection bias.
  • Control for time-invariant confounders.
  • Requires a rigorous research design.
  • Sensitive to measurement errors.
  • Difficult to interpret when multiple interventions or events are happening during the study period.

Challenges and Best Practices

  • Validity Threat
  • Selection bias : This occurs when the groups being compared are not equivalent. This can happen if participants are self-selected into the groups, or if the groups are not randomly assigned.
  • History effects : These are events that happen during the study period that could affect the dependent variable. For example, if there is a natural disaster during the study period, it could affect the results.
  • Maturation : This refers to the natural changes that occur over time. For example, students may naturally improve their test scores over time, regardless of whether or not they receive an intervention.
  • Testing effects : These are the effects of taking a test on subsequent test scores. For example, if students take a pretest before an intervention, they may learn from the test and do better on the posttest, even if the intervention had no effect.
  • Instrumentation : This refers to changes in the way the dependent variable is measured. For example, if the test used to measure student achievement is changed during the study period, it could affect the results.

Strategies for Minimizing Validity Threats

  • Matching participants on important characteristics such as age, education level, and more to reduce selection bias.
  • Use control groups to avoid history effects, maturation effects, and testing effects.
  • Use multiple methods to measure dependent variables to reduce instrumentation effects.
  • Conduct a pilot study to identify and address potential validity threats 

Read Also – Internal Validity in Research: Definition, Threats, Examples 

Sample Size and Power

Sample size is the number of participants in a study, while power is the probability of detecting a meaningful effect if it exists.

You have to carefully consider sample size and power when designing a quasi-experiment. This is because groups may not be 100% homogeneous at the start of the study, which can limit the strength of the design.

Using power analysis, you can figure out the sample size you need to see a significant effect. Power analysis looks at the magnitude of the effect, and the standard deviation of the dependent variable, and determines the alpha level.

Generalizability

A major downside of the quasi-experimental design is that it’s usually not generalizable or applicable to other environments. It is typically done in natural environments, so you can’t control factors that could influence the results.

Carefully consider the context of the study before you generalize a quasi-experimental. Also, try replicating the study in other settings to see if the results are consistent.

Real-World Applications

  • Healthcare Interventions:

A study by the Universiti Kebangsaan, Malaysia used a quasi-experimental design to assess the effectiveness of a new program for preventing childhood obesity. The study found that the program was effective in reducing the risk of obesity, but it was also expensive.

  • Education Policy Changes

A study by Raj Chetty and his colleagues found that students who attended charter schools in California were more likely to attend college than students who did not attend charter schools. However, this study arguably promoted academically underqualified students being admitted to colleges.

  • Social and Economic Interventions

A study by the RAND Corporation used a quasi-experimental design to assess the effects of a job training program on employment and earnings. 

The study found that job training programs were effective in increasing employment and earnings, but they also found that the impact varied depending on the characteristics of the participants and the design of the program.

Ethical Considerations

  • Informed consent : Provide full information about the purpose of the study, the procedures, the risks and benefits of participating, and their right to withdraw from the study at any time.
  • Confidentiality : Do not collect any personal information that is irrelevant to the study and keep participant information confidential.
  • Risks to participants : Quasi-experimental research may pose some risks to participants, such as the risk of harm or discomfort. Minimize these risks as much as possible and only perform the research if the benefits outweigh the risks.
  • Benefits to participants : Quasi-experimental research should offer some potential benefits to participants, such as access to new treatments or interventions. Researchers should carefully consider the potential benefits and risks of participating in the study before recruiting participants
  • Balance of research goals and participant welfare : Do not conduct research that is likely to harm participants, even if the research has the potential to benefit society.

Read: What are Ethical Practices in Market Research?

Quasi-experimental research is a valuable tool for understanding the causal effects of interventions. It is particularly useful when you can’t conduct actual experiments because of ethical or practical constraints.

However, it is important to be aware of the limitations of this type of research. Carefully design the study and consider the limitations to ensure that the findings are accurate and reliable.

Logo

Connect to Formplus, Get Started Now - It's Free!

  • Quasi-Experimental Research
  • Quasi-Experiments
  • True Experiments
  • Moradeke Owa

Formplus

You may also like:

Market Research: Types, Methods & Survey Examples

A complete guide on market research; definitions, survey examples, templates, importance and tips.

quasi experimental definition in research

What is Pure or Basic Research? + [Examples & Method]

Simple guide on pure or basic research, its methods, characteristics, advantages, and examples in science, medicine, education and psychology

Recall Bias: Definition, Types, Examples & Mitigation

This article will discuss the impact of recall bias in studies and the best ways to avoid them during research.

Exploratory Research: What are its Method & Examples?

Overview on exploratory research, examples and methodology. Shows guides on how to conduct exploratory research with online surveys

Formplus - For Seamless Data Collection

Collect data the right way with a versatile data collection tool. try formplus and transform your work productivity today..

Instant insights, infinite possibilities

The use and interpretation of quasi-experimental design

Last updated

6 February 2023

Reviewed by

Miroslav Damyanov

Short on time? Get an AI generated summary of this article instead

  • What is a quasi-experimental design?

Commonly used in medical informatics (a field that uses digital information to ensure better patient care), researchers generally use this design to evaluate the effectiveness of a treatment – perhaps a type of antibiotic or psychotherapy, or an educational or policy intervention.

Even though quasi-experimental design has been used for some time, relatively little is known about it. Read on to learn the ins and outs of this research design.

Make research less tedious

Dovetail streamlines research to help you uncover and share actionable insights

  • When to use a quasi-experimental design

A quasi-experimental design is used when it's not logistically feasible or ethical to conduct randomized, controlled trials. As its name suggests, a quasi-experimental design is almost a true experiment. However, researchers don't randomly select elements or participants in this type of research.

Researchers prefer to apply quasi-experimental design when there are ethical or practical concerns. Let's look at these two reasons more closely.

Ethical reasons

In some situations, the use of randomly assigned elements can be unethical. For instance, providing public healthcare to one group and withholding it to another in research is unethical. A quasi-experimental design would examine the relationship between these two groups to avoid physical danger.

Practical reasons

Randomized controlled trials may not be the best approach in research. For instance, it's impractical to trawl through large sample sizes of participants without using a particular attribute to guide your data collection .

Recruiting participants and properly designing a data-collection attribute to make the research a true experiment requires a lot of time and effort, and can be expensive if you don’t have a large funding stream.

A quasi-experimental design allows researchers to take advantage of previously collected data and use it in their study.

  • Examples of quasi-experimental designs

Quasi-experimental research design is common in medical research, but any researcher can use it for research that raises practical and ethical concerns. Here are a few examples of quasi-experimental designs used by different researchers:

Example 1: Determining the effectiveness of math apps in supplementing math classes

A school wanted to supplement its math classes with a math app. To select the best app, the school decided to conduct demo tests on two apps before selecting the one they will purchase.

Scope of the research

Since every grade had two math teachers, each teacher used one of the two apps for three months. They then gave the students the same math exams and compared the results to determine which app was most effective.

Reasons why this is a quasi-experimental study

This simple study is a quasi-experiment since the school didn't randomly assign its students to the applications. They used a pre-existing class structure to conduct the study since it was impractical to randomly assign the students to each app.

Example 2: Determining the effectiveness of teaching modern leadership techniques in start-up businesses

A hypothetical quasi-experimental study was conducted in an economically developing country in a mid-sized city.

Five start-ups in the textile industry and five in the tech industry participated in the study. The leaders attended a six-week workshop on leadership style, team management, and employee motivation.

After a year, the researchers assessed the performance of each start-up company to determine growth. The results indicated that the tech start-ups were further along in their growth than the textile companies.

The basis of quasi-experimental research is a non-randomized subject-selection process. This study didn't use specific aspects to determine which start-up companies should participate. Therefore, the results may seem straightforward, but several aspects may determine the growth of a specific company, apart from the variables used by the researchers.

Example 3: A study to determine the effects of policy reforms and of luring foreign investment on small businesses in two mid-size cities

In a study to determine the economic impact of government reforms in an economically developing country, the government decided to test whether creating reforms directed at small businesses or luring foreign investments would spur the most economic development.

The government selected two cities with similar population demographics and sizes. In one of the cities, they implemented specific policies that would directly impact small businesses, and in the other, they implemented policies to attract foreign investment.

After five years, they collected end-of-year economic growth data from both cities. They looked at elements like local GDP growth, unemployment rates, and housing sales.

The study used a non-randomized selection process to determine which city would participate in the research. Researchers left out certain variables that would play a crucial role in determining the growth of each city. They used pre-existing groups of people based on research conducted in each city, rather than random groups.

  • Advantages of a quasi-experimental design

Some advantages of quasi-experimental designs are:

Researchers can manipulate variables to help them meet their study objectives.

It offers high external validity, making it suitable for real-world applications, specifically in social science experiments.

Integrating this methodology into other research designs is easier, especially in true experimental research. This cuts down on the time needed to determine your outcomes.

  • Disadvantages of a quasi-experimental design

Despite the pros that come with a quasi-experimental design, there are several disadvantages associated with it, including the following:

It has a lower internal validity since researchers do not have full control over the comparison and intervention groups or between time periods because of differences in characteristics in people, places, or time involved. It may be challenging to determine whether all variables have been used or whether those used in the research impacted the results.

There is the risk of inaccurate data since the research design borrows information from other studies.

There is the possibility of bias since researchers select baseline elements and eligibility.

  • What are the different quasi-experimental study designs?

There are three distinct types of quasi-experimental designs:

Nonequivalent

Regression discontinuity, natural experiment.

This is a hybrid of experimental and quasi-experimental methods and is used to leverage the best qualities of the two. Like the true experiment design, nonequivalent group design uses pre-existing groups believed to be comparable. However, it doesn't use randomization, the lack of which is a crucial element for quasi-experimental design.

Researchers usually ensure that no confounding variables impact them throughout the grouping process. This makes the groupings more comparable.

Example of a nonequivalent group design

A small study was conducted to determine whether after-school programs result in better grades. Researchers randomly selected two groups of students: one to implement the new program, the other not to. They then compared the results of the two groups.

This type of quasi-experimental research design calculates the impact of a specific treatment or intervention. It uses a criterion known as "cutoff" that assigns treatment according to eligibility.

Researchers often assign participants above the cutoff to the treatment group. This puts a negligible distinction between the two groups (treatment group and control group).

Example of regression discontinuity

Students must achieve a minimum score to be enrolled in specific US high schools. Since the cutoff score used to determine eligibility for enrollment is arbitrary, researchers can assume that the disparity between students who only just fail to achieve the cutoff point and those who barely pass is a small margin and is due to the difference in the schools that these students attend.

Researchers can then examine the long-term effects of these two groups of kids to determine the effect of attending certain schools. This information can be applied to increase the chances of students being enrolled in these high schools.

This research design is common in laboratory and field experiments where researchers control target subjects by assigning them to different groups. Researchers randomly assign subjects to a treatment group using nature or an external event or situation.

However, even with random assignment, this research design cannot be called a true experiment since nature aspects are observational. Researchers can also exploit these aspects despite having no control over the independent variables.

Example of the natural experiment approach

An example of a natural experiment is the 2008 Oregon Health Study.

Oregon intended to allow more low-income people to participate in Medicaid.

Since they couldn't afford to cover every person who qualified for the program, the state used a random lottery to allocate program slots.

Researchers assessed the program's effectiveness by assigning the selected subjects to a randomly assigned treatment group, while those that didn't win the lottery were considered the control group.

  • Differences between quasi-experiments and true experiments

There are several differences between a quasi-experiment and a true experiment:

Participants in true experiments are randomly assigned to the treatment or control group, while participants in a quasi-experiment are not assigned randomly.

In a quasi-experimental design, the control and treatment groups differ in unknown or unknowable ways, apart from the experimental treatments that are carried out. Therefore, the researcher should try as much as possible to control these differences.

Quasi-experimental designs have several "competing hypotheses," which compete with experimental manipulation to explain the observed results.

Quasi-experiments tend to have lower internal validity (the degree of confidence in the research outcomes) than true experiments, but they may offer higher external validity (whether findings can be extended to other contexts) as they involve real-world interventions instead of controlled interventions in artificial laboratory settings.

Despite the distinct difference between true and quasi-experimental research designs, these two research methodologies share the following aspects:

Both study methods subject participants to some form of treatment or conditions.

Researchers have the freedom to measure some of the outcomes of interest.

Researchers can test whether the differences in the outcomes are associated with the treatment.

  • An example comparing a true experiment and quasi-experiment

Imagine you wanted to study the effects of junk food on obese people. Here's how you would do this as a true experiment and a quasi-experiment:

How to carry out a true experiment

In a true experiment, some participants would eat junk foods, while the rest would be in the control group, adhering to a regular diet. At the end of the study, you would record the health and discomfort of each group.

This kind of experiment would raise ethical concerns since the participants assigned to the treatment group are required to eat junk food against their will throughout the experiment. This calls for a quasi-experimental design.

How to carry out a quasi-experiment

In quasi-experimental research, you would start by finding out which participants want to try junk food and which prefer to stick to a regular diet. This allows you to assign these two groups based on subject choice.

In this case, you didn't assign participants to a particular group, so you can confidently use the results from the study.

When is a quasi-experimental design used?

Quasi-experimental designs are used when researchers don’t want to use randomization when evaluating their intervention.

What are the characteristics of quasi-experimental designs?

Some of the characteristics of a quasi-experimental design are:

Researchers don't randomly assign participants into groups, but study their existing characteristics and assign them accordingly.

Researchers study the participants in pre- and post-testing to determine the progress of the groups.

Quasi-experimental design is ethical since it doesn’t involve offering or withholding treatment at random.

Quasi-experimental design encompasses a broad range of non-randomized intervention studies. This design is employed when it is not ethical or logistically feasible to conduct randomized controlled trials. Researchers typically employ it when evaluating policy or educational interventions, or in medical or therapy scenarios.

How do you analyze data in a quasi-experimental design?

You can use two-group tests, time-series analysis, and regression analysis to analyze data in a quasi-experiment design. Each option has specific assumptions, strengths, limitations, and data requirements.

Should you be using a customer insights hub?

Do you want to discover previous research faster?

Do you share your research findings with others?

Do you analyze research data?

Start for free today, add your research, and get to key insights faster

Editor’s picks

Last updated: 18 April 2023

Last updated: 27 February 2023

Last updated: 22 August 2024

Last updated: 5 February 2023

Last updated: 16 April 2023

Last updated: 9 March 2023

Last updated: 30 April 2024

Last updated: 12 December 2023

Last updated: 11 March 2024

Last updated: 4 July 2024

Last updated: 6 March 2024

Last updated: 5 March 2024

Last updated: 13 May 2024

Latest articles

Related topics, .css-je19u9{-webkit-align-items:flex-end;-webkit-box-align:flex-end;-ms-flex-align:flex-end;align-items:flex-end;display:-webkit-box;display:-webkit-flex;display:-ms-flexbox;display:flex;-webkit-flex-direction:row;-ms-flex-direction:row;flex-direction:row;-webkit-box-flex-wrap:wrap;-webkit-flex-wrap:wrap;-ms-flex-wrap:wrap;flex-wrap:wrap;-webkit-box-pack:center;-ms-flex-pack:center;-webkit-justify-content:center;justify-content:center;row-gap:0;text-align:center;max-width:671px;}@media (max-width: 1079px){.css-je19u9{max-width:400px;}.css-je19u9>span{white-space:pre;}}@media (max-width: 799px){.css-je19u9{max-width:400px;}.css-je19u9>span{white-space:pre;}} decide what to .css-1kiodld{max-height:56px;display:-webkit-box;display:-webkit-flex;display:-ms-flexbox;display:flex;-webkit-align-items:center;-webkit-box-align:center;-ms-flex-align:center;align-items:center;}@media (max-width: 1079px){.css-1kiodld{display:none;}} build next, decide what to build next, log in or sign up.

Get started for free

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

quasi experimental definition in research

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:

 Experimental Study (a.k.a. Randomized Controlled Trial)Quasi-Experimental Study
ObjectiveEvaluate the effect of an intervention or a treatmentEvaluate the effect of an intervention or a treatment
How participants get assigned to groups?Random assignmentNon-random assignment (participants get assigned according to their choosing or that of the researcher)
Is there a control group?YesNot always (although, if present, a control group will provide better evidence for the study results)
Is there any room for confounding?No (although check for a detailed discussion on post-randomization confounding in randomized controlled trials)Yes (however, statistical techniques can be used to study causal relationships in quasi-experiments)
Level of evidenceA randomized trial is at the highest level in the hierarchy of evidenceA quasi-experiment is one level below the experimental study in the hierarchy of evidence [ ]
AdvantagesMinimizes bias and confounding– Can be used in situations where an experiment is not ethically or practically feasible
– Can work with smaller sample sizes than randomized trials
Limitations– High cost (as it generally requires a large sample size)
– Ethical limitations
– Generalizability issues
– Sometimes practically infeasible
Lower ranking in the hierarchy of evidence as losing the power of randomization causes the study to be more susceptible to bias and confounding

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
  • Skip to main content
  • Skip to primary sidebar

IResearchNet

Quasi-Experimental Design

Quasi-experimental design definition.

Quasi-Experimental Design

Example of a Quasi-Experimental Design

Quasi-experimental designs are most often used in natural (nonlaboratory) settings over longer periods and usually include an intervention or treatment. Consider, for example, a study of the effect of a motivation intervention on class attendance and enjoyment in students. When an intact group such as a classroom is singled out for an intervention, randomly assigning each person to experimental conditions is not possible. Rather, the researcher gives one classroom the motivational intervention (intervention group) and the other classroom receives no intervention (comparison group). The researcher uses two classrooms that are as similar as possible in background (e.g., same age, racial composition) and that have comparable experiences within the class (e.g., type of class, meeting time) except for the intervention. In addition, the researcher gives participants in both conditions (comparison and motivation intervention) pretest questionnaires to assess attendance, enjoyment, and other related variables before the intervention. After the intervention is administered, the researcher measures attendance and enjoyment of the class. The researcher can then determine if students in the motivation intervention group enjoyed and attended class more than the students in the comparison group did.

Interpreting Results from a Quasi-Experimental Design

How should results from this hypothetical study be interpreted? Investigators, when interpreting the results of quasi-experimental designs that lacked random assignment of participants to conditions, must be cautious drawing conclusions about causality because of potential confounds in the setting. For example, the previous hypothetical example course material in the intervention group might have become more engaging whereas the comparison group started to cover a more mundane topic that led to changes in class enjoyment and attendance. However, if the intervention group and comparison group had similar pretest scores and comparable classroom experiences, then changes on posttest scores suggest that the motivation intervention influenced class attendance and enjoyment.

The Pros and Cons of Using Quasi-Experimental Designs

Quasi-experiments are most useful when conducting research in settings where random assignment is not possible because of ethical considerations or constraining situational factors. In consequence, such designs are more prevalent in studies conducted in natural settings, thereby increasing the real-world applicability of the findings. Such studies are not, however, true experiments, and thus the lack of control over assignment of participants to conditions renders causal conclusions suspect.

References:

  • Cook, T. D., & Campbell, D. T. (1979). Quasi-experimental: Design and analysis issues for field settings. Boston: Houghton Mifflin.
  • Shadish, W. R., Cook, T. D., & Campbell, D. T. (2002). Experimental and quasi-experimental designs for generalized causal inference. Boston: Houghton Mifflin.
  • Social Psychology Research Methods

U.S. flag

An official website of the United States government

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

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

  • Publications
  • Account settings

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

  • Advanced Search
  • Journal List
  • HHS Author Manuscripts

Logo of nihpa

Quasi-Experimental Designs for Causal Inference

When randomized experiments are infeasible, quasi-experimental designs can be exploited to evaluate causal treatment effects. The strongest quasi-experimental designs for causal inference are regression discontinuity designs, instrumental variable designs, matching and propensity score designs, and comparative interrupted time series designs. This article introduces for each design the basic rationale, discusses the assumptions required for identifying a causal effect, outlines methods for estimating the effect, and highlights potential validity threats and strategies for dealing with them. Causal estimands and identification results are formalized with the potential outcomes notations of the Rubin causal model.

Causal inference plays a central role in many social and behavioral sciences, including psychology and education. But drawing valid causal conclusions is challenging because they are warranted only if the study design meets a set of strong and frequently untestable assumptions. Thus, studies aiming at causal inference should employ designs and design elements that are able to rule out most plausible threats to validity. Randomized controlled trials (RCTs) are considered as the gold standard for causal inference because they rely on the fewest and weakest assumptions. 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 score designs, and the comparative interrupted time series design. For each design we outline the strategy and assumptions for identifying a causal effect, address estimation methods, and discuss practical issues and suggestions for strengthening the basic designs. To highlight the design differences, throughout the article we use a hypothetical example with the following causal research question: What is the effect of attending a summer science camp on students’ science achievement?

POTENTIAL OUTCOMES AND RANDOMIZED CONTROLLED TRIAL

Before we discuss the four quasi-experimental designs, we introduce the potential outcomes notation of the Rubin causal model (RCM) and show how it is used in the context of an RCT. The RCM ( Holland, 1986 ) formalizes causal inference in terms of potential outcomes, which allow us to precisely define causal quantities of interest and to explicate the assumptions required for identifying them. RCM considers a potential outcome for each possible treatment condition. For a dichotomous treatment variable (i.e., a treatment and control condition), each subject i has a potential treatment outcome Y i (1), which we would observe if subject i receives the treatment ( Z i = 1), and a potential control outcome Y i (0), which we would observe if subject i receives the control condition ( Z i = 0). The difference in the two potential outcomes, Y i (1)− Y i (0), represents the individual causal effect.

Suppose we want to evaluate the effect of attending a summer science camp on students’ science achievement score. Then each student has two potential outcomes: a potential control score for not attending the science camp, and the potential treatment score for attending the camp. However, the individual causal effects of attending the camp cannot be inferred from data, because the two potential outcomes are never observed simultaneously. Instead, researchers typically focus on average causal effects. The average treatment effect (ATE) for the entire study population is defined as the difference in the expected potential outcomes, ATE = E [ Y i (1)] − E [ Y i (0)]. Similarly, we can also define the ATE for the treated subjects (ATT), ATT = E [ Y i (1) | Z i = 1] − E [ Y (0) | Z i =1]. Although the expectations of the potential outcomes are not directly observable because not all potential outcomes are observed, we nonetheless can identify ATE or ATT under some reasonable assumptions. In an RCT, random assignment establishes independence between the potential outcomes and the treatment status, which allows us to infer ATE. Suppose that students are randomly assigned to the science camp and that all students comply with the assigned condition. Then random assignment guarantees that the camp attendance indicator Z is independent of the potential achievement scores Y i (0) and Y i (1).

The independence assumption allows us to rewrite ATE in terms of observable expectations (i.e., with observed outcomes instead of potential outcomes). First, due to the independence (randomization), the unconditional expectations of the potential outcomes can be expressed as conditional expectations, E [ Y i (1)] = E [ Y i (1) | Z i = 1] and E [ Y i (0)] = E [ Y i (0) | Z i = 0] Second, because the potential treatment outcomes are actually observed for the treated, we can replace the potential treatment outcome with the observed outcome such that E [ Y i (1) | Z i = 1] = E [ Y i | Z i = 1] and, analogously, E [ Y i (0) | Z i = 0] = E [ Y i | Z i = 0] Thus, the ATE is expressible in terms of observable quantities rather than potential outcomes, ATE = E [ Y i (1)] − E [ Y i (0)] = E [ Y i | Z i = 1] – E [ Y i | Z i = 0], and we that say ATE is identified.

This derivation also rests on the stable-unit-treatment-value assumption (SUTVA; Imbens & Rubin, 2015 ). SUTVA is required to properly define the potential outcomes, that is, (a) the potential outcomes of a subject depend neither on the assignment mode nor on other subjects’ treatment assignment, and (b) there is only one unique treatment and one unique control condition. Without further mentioning, we assume SUTVA for all quasi-experimental designs discussed in this article.

REGRESSION DISCONTINUITY DESIGN

Due to ethical or budgetary reasons, random assignment is often infeasible in practice. Nonetheless, researchers may sometimes still retain full control over treatment assignment as in a regression discontinuity (RD) design where, based on a continuous assignment variable and a cutoff score, subjects are deterministically assigned to treatment conditions.

Suppose that the science camp is a remedial program and only students whose grade point average (GPA) score is less than or equal to 2.0 are eligible to participate. Figure 1 shows a scatterplot of hypothetical data where the x-axis represents the assignment variable ( GPA ) and the y -axis the outcome ( Science Score ). All subjects with a GPA score below the cutoff attended the camp (circles), whereas all subjects scoring above the cutoff do not attend (squares). Because all low-achieving students are in the treatment group and all high-achieving students in the control group, their respective GPA distributions do not overlap, not even at the cutoff. This lack of overlap complicates the identification of a causal effect because students in the treatment and control group are not comparable at all (i.e., they have a completely different distribution of the GPA scores).

An external file that holds a picture, illustration, etc.
Object name is nihms-983980-f0001.jpg

A hypothetical example of regression discontinuity design. Note . GPA = grade point average.

One strategy of dealing with the lack of overlap is to rely on the linearity assumption of regression models and to extrapolate into areas of nonoverlap. However, if the linear models do not correctly specify the functional form, the resulting ATE estimate is biased. A safer strategy is to evaluate the treatment effect only at the cutoff score where treatment and control cases almost overlap, and thus functional form assumptions and extrapolation are almost no longer needed. Consider the treatment and control students that score right at the cutoff or just above it. Students with a GPA score of 2.0 participate in the science camp and students with a GPA score of 2.1 are in the control condition (the status quo condition or a different camp). The two groups of students are essentially equivalent because the difference in their GPA scores is negligibly small (2.1 − 2.0 = .1) and likely due to random chance (measurement error) rather than a real difference in ability. Thus, in the very close neighborhood around the cutoff score, the RD design is equivalent to an RCT; therefore, the ATE at the cutoff (ATEC) is identified.

CAUSAL ESTIMAND AND IDENTIFICATION

ATEC is defined as the difference in the expected potential treatment and control outcomes for the subjects scoring exactly at the cutoff: ATEC = E [ Y i (1) | A i = a c ] − E [ Y i (0) | A i = a c ], where A denotes assignment variable and a c the cutoff score. Because we observe only treatment subjects and not control subjects right at the cutoff, we need two assumptions in order to identify ATEC ( Hahn, Todd, & van Klaauw, 2001 ): (a) the conditional expectations of the potential treatment and control outcomes are continuous at the cutoff ( continuity ), and (b) all subjects comply with treatment assignment ( full compliance ).

The continuity assumption can be expressed in terms of limits as lim a ↓ a C E [ Y i ( 1 ) | A i = a ] = E [ Y i ( 1 ) | A i = a ] = lim a ↑ a C E [ Y i ( 1 ) | A i = a ] and lim a ↓ a C E [ Y i ( 0 ) | A i = a ] = E [ Y i ( 0 ) | A i = a ] = lim a ↑ a C E [ Y i ( 0 ) | A i = a ] . Thus, we can rewrite ATEC as the difference in limits, A T E C = lim a ↑ a C E [ Y i ( 1 ) | A i = a c ] − lim a ↓ a C E [ Y i ( 0 ) | A i = a c ] , which solves the issue that no control subjects are observed directly at the cutoff. Then, by the full compliance assumption, the potential treatment and control outcomes can be replaced with the observed outcomes such that A T E C = lim a ↑ a C E [ Y i | A i = a c ] − lim a ↓ a C E [ Y i | A i = a c ] is identified at the cutoff (i.e., ATEC is now expressed in terms of observable quantities). The difference in the limits represents the discontinuity in the mean outcomes exactly at the cutoff ( Figure 1 ).

Estimating ATEC

ATEC can be estimated with parametric or nonparametric regression methods. First, consider the parametric regression of the outcome Y on the treatment Z , the cutoff-centered assignment variable A − a c , and their interaction: Y = β 0 + β 1 Z + β 2 ( A − a c ) + β 3 ( Z × ( A − a c )) + e . If the model correctly specifies the functional form, then β ^ 1 is an unbiased estimator for ATEC. In practice, an appropriate model specification frequently involves also quadratic and cubic terms of the assignment variable plus their interactions with the treatment indicator.

To avoid overly strong functional form assumptions, semiparametric or nonparametric regression methods like generalized additive models or local linear kernel regression can be employed ( Imbens & Lemieux, 2008 ). These methods down-weight or even discard observations that are not in the close neighborhood around the cutoff. The R packages rdd ( Dimmery, 2013 ) and rdrobust ( Calonico, Cattaneo, & Titiunik, 2015 ), or the command rd in STATA ( Nichols, 2007 ) are useful for estimation and diagnostic purposes.

Practical Issues

A major validity threat for RD designs is the manipulation of the assignment score around the cutoff, which directly results in a violation of the continuity assumption ( Wong et al., 2012 ). For instance, if a teacher knows the assignment score in advance and he wants all his students to attend the science camp, the teacher could falsely report a GPA score of 2.0 or below for the students whose actual GPA score exceeds the cutoff value.

Another validity threat is noncompliance, meaning that subjects assigned to the control condition may cross over to the treatment and subjects assigned to the treatment do not show up. An RD design with noncompliance is called a fuzzy RD design (instead of a sharp RD design with full compliance). A fuzzy RD design still allows us to identify the intention-to-treat effect or the local average treatment effect at the cutoff (LATEC). The intention-to-treat effect refers to the effect of treatment assignment rather than the actual treatment receipt. LATEC estimates ATEC for the subjects who comply with treatment assignment. LATEC is identified if one uses the assignment status as an instrumental variable for treatment receipt (see the upcoming Instrumental Variable section).

Finally, generalizability and statistical power are often mentioned as major disadvantages of RD designs. Because RD designs identify the treatment effect only at the cutoff, ATEC estimates are not automatically generalizable to subjects scoring further away from the cutoff. Statistical power for detecting a significant effect is an issue because the lack of overlap on the assignment variable results in increased standard errors. With semi- or nonparametric regression methods, power further diminishes.

Strengthening RD Designs

To avoid systematic manipulations of the assignment variable, it is desirable to conceal the assignment rule from study participants and administrators. If the assignment rule is known to them, manipulations can hardly be ruled out, particularly when the stakes are high. Researchers can use the McCrary test ( McCrary, 2008 ) to check for potential manipulations. The test investigates whether there is a discontinuity in the distribution of the assignment variable right at the cutoff. Plotting baseline covariates against the assignment variable, and regressing the covariates on the assignment variable and the treatment indicator also help in detecting potential discontinuities at the cutoff.

The RD design’s validity can be increased by combining the basic RD design with other designs. An example is the tie-breaking RD design, which uses two cutoff scores. Subjects scoring between the two cutoff scores are randomly assigned to treatment conditions, whereas subjects scoring outside the cutoff interval receive the treatment or control condition according to the RD assignment rule ( Black, Galdo & Smith, 2007 ). This design combines an RD design with an RCT and is advantageous with respect to the correct specification of the functional form, generalizability, and statistical power. Similar benefits can be obtained by adding pretest measures of the outcome or nonequivalent comparison groups ( Wing & Cook, 2013 ).

Imbens and Lemieux (2008) and Lee and Lemieux (2010) provided comprehensive introductions to RD designs. Lee and Lemieux also summarized many applications from economics. Angrist and Lavy (1999) applied the design to investigate the effect of class size on student achievement.

INSTRUMENTAL VARIABLE DESIGN

In practice, researchers often have no or only partial control over treatment selection. In addition, they might also lack reliable knowledge of the selection process. Nonetheless, even with limited control and knowledge of the selection process it is still possible to identify a causal treatment effect if an instrumental variable (IV) is available. An IV is an exogenous variable that is related to the treatment but is completely unrelated to the outcome, except via treatment. An IV design requires researchers either to create an IV at the design stage (as in an encouragement design; see next) or to find an IV in the data set at hand or a related data base.

Consider the science camp example, but instead of random or deterministic treatment assignment, students decide on their own or together with their parents whether to attend the camp. Many factors may determine the decision, for instance, students’ science ability and motivation, parents’ socioeconomic status, or the availability of public transportation for the daily commute to the camp. Whereas the first three variables are presumably also related to the science outcome, public transportation might be unrelated to the science score (except via camp attendance). Thus, the availability of public transportation may qualify as an IV. Figure 2 illustrates such IV design: Public transportation (IV) directly affects camp attendance but has no direct or indirect effect on science achievement (outcome) other than through camp attendance (treatment). The question mark represents unknown or unobserved confounders, that is, variables that simultaneously affect both camp attendance and science achievement. The IV design allows us to identify a causal effect even if some or all confounders are unknown or unobserved.

An external file that holds a picture, illustration, etc.
Object name is nihms-983980-f0002.jpg

A diagram of an example of instrumental variable design.

The strategy for identifying a causal effect is based on exploiting the variation in the treatment variable explained by IV. In Figure 2 , the total variation in the treatment consists of (a) the variation induced by the IV and (b) the variation induced by confounders (question mark) and other exogenous variables (not shown in the figure). The identification of the camp’s effect requires us to isolate the treatment variation that is related to public transportation (IV), and then to use the isolated variation to investigate the camp’s effect on the science score. Because we exploit the treatment variation exclusively induced by the IV but ignore the variation induced by unobserved or unknown confounders, the IV design identifies the ATE for the sub-population of compliers only. In our example, the compliers are the students who attend the camp because public transportation is available and do not attend because it is unavailable. For students whose parents always use their own car to drop them off and pick them up at the camp location, we cannot infer the causal effect, because their camp attendance is completely unrelated to the availability of public transportation.

Causal Estimand and Identification

The complier average treatment effect (CATE) is defined as the expected difference in potential outcomes for the sub-population of compliers: CATE = E [ Y i (1) | Complier ] − E [ Y i (0) | Complier ] = τ C .

Identification requires us to distinguish between four latent groups: compliers (C), who attend the camp if public transportation is available but do not attend if unavailable; always-takers (A), who always attend the camp regardless of whether or not public transportation is available; never-takers (N), who never attend the camp regardless of public transportation; and defiers (D), who do not attend if public transportation is available but attend if unavailable. Because group membership is unknown, it is impossible to directly infer CATE from the data of compliers. However, CATE is identified from the entire data set if (a) the IV is predictive of the treatment ( predictive first stage ), (b) the IV is unrelated to the outcome except via treatment ( exclusion restriction ), and (c) no defiers are present ( monotonicity ; Angrist, Imbens, & Rubin, 1996 ; see Steiner, Kim, Hall, & Su, 2015 , for a graphical explanation).

First, notice that the IV’s effects on the treatment (γ) and the outcome (δ) are directly identified from the observed data because the IV’s relation with the treatment and outcome is unconfounded. In our example ( Figure 2 ), γ denotes the effect of public transportation on camp attendance and δ the indirect effect of public transportation on the science score. Both effects can be written as weighted averages of the corresponding group-specific effects ( γ C , γ A , γ N , γ D and δ C , δ A , δ N , δ D for compliers, always-takers, never-takers, and defiers, respectively): γ = p ( C ) γ C + p ( A ) γA + p ( N ) γ N + p ( D ) γ D and δ = p ( C ) δ C + p ( A ) δ A + p ( N ) δ N + p ( D ) δ D where p (.) represents the portion of the respective latent group in the population and p ( C ) + p ( A ) + p ( N ) + p ( D ) = 1. Because the treatment choice of always-takers and never-takers is entirely unaffected by the instrument, the IV’s effect on the treatment is zero, γ A = γ N = .0, and together with the exclusion restriction , we also know δ A = δ N = 0, that is, the IV has no effect on the outcome. If no defiers are present, p ( D ) = 0 ( monotonicity ), then the IV’s effects on the treatment and outcome simplify to γ = p ( C ) γC and δ = p ( C ) δC , respectively. Because δ C = γ C τ C and γ ≠ 0 ( predictive first stage ), the ratio of the observable IV effects, γ and δ, identifies CATE: δ γ = p ( C ) γ C τ C p ( C ) γ C = τ C .

Estimating CATE

A two-stage least squares (2SLS) regression is typically used for estimating CATE. In the first stage, treatment Z is regressed on the IV, Z = β 0 + β 1 IV + e . The linear first-stage model applies with a dichotomous treatment variable (linear probability model). The second stage then regresses the outcome Y on the predicted values Z ^ from the first stage model, Y = π 0 + π 1 Z ^ + r , where π ^ 1 is the CATE estimator. The two stages are automatically performed by the 2SLS procedure, which also provides an appropriate standard error for the effect estimate. The STATA commands ivregress and ivreg2 ( Baum, Schaffer, & Stillman, 2007 ) or the sem package in R ( Fox, 2006 ) perform the 2SLS regression.

One challenge in implementing an IV design is to find a valid instrument that satisfies the assumptions just discussed. In particular, the exclusion restriction is untestable and frequently hard to defend in practice. In our example, if high-income families live in suburban areas with bad public transportation connections, then the availability of the public transportation is likely related to the science score via household income (or socioeconomic status). Although conditioning on the observed household income can transform public transportation into a conditional IV (see next), one can frequently come up with additional scenarios that explains why the IV is related to the outcome and thus violates the exclusion restriction.

Another issue arises from “weak” IVs that are only weakly related to treatment. Weak IVs cause efficiency problems ( Wooldridge, 2012 ). If the availability of public transportation barely affects camp attendance because most parents give their children a ride anyway, the IV’s effect on the treatment ( γ ) is close to zero. Because γ ^ is the denominator in the CATE estimator, τ ^ C = δ ^ / γ ^ , an imprecisely estimated γ ^ results in a considerable over- or underestimation of CATE. Moreover, standard errors will be large.

One also needs to keep in mind that the substantive meaning of CATE depends on the chosen IV. Consider two slightly different IVs with respect to public transportation: the availability of (a) a bus service and (b) subway service. For the first IV, the complier population consists of students who choose to (not) attend the camp depending on the availability of a bus service. For the second IV, the complier population refers to the availability of a subway service. Because the two complier populations are very likely different from each other (students who are willing to take the subway might not be willing to take the bus), the corresponding CATEs refer to different subpopulations.

Strengthening IV Designs

Given the challenges in identifying a valid instrument from observed data, researchers should consider creating an IV at the design stage of a study. Although it might be impossible to directly assign subjects to treatment conditions, one might still be able to encourage participants to take the treatment. Subjects are randomly encouraged to sign up for treatment, but whether they actually comply with the encouragement is entirely their own decision ( Imai et al., 2011 ). Random encouragement qualifies as an IV because it very likely meets the exclusion restriction. For example, instead of collecting data on public transportation, researchers may advertise and recommend the science camp in a letter to the parents of a randomly selected sample of students.

With observational data it is hard to identify a valid IV because covariates that strongly predict the treatment are usually also related to the outcome. However, these covariates can still qualify as an IV if they affect the outcome only indirectly via other observed variables. Such covariates can be used as conditional IVs, that is, they meet the IV requirements conditional on the observed variables ( Brito & Pearl, 2002 ). Assume the availability of public transportation (IV) is associated with the science score via household income. Then, controlling for the reliably measured household income in both stages of the 2SLS analysis blocks the IV’s relation to the science score and turns public transportation into a conditional IV. However, controlling for a large set of variables does not guarantee that the exclusion restriction is more likely met. It may even result in more bias as compared to an IV analysis with fewer covariates ( Ding & Miratrix, 2015 ; Steiner & Kim, in press ). The choice of a valid conditional IV requires researchers to carefully select the control variables based on subject-matter theory.

The seminal article by Angrist et al. (1996) provides a thorough discussion of the IV design, and Steiner, Kim, et al. (2015 ) proved the identification result using graphical models. Excellent introductions to IV designs can be found in Angrist and Pischke (2009 , 2015) . Angrist and Krueger (1992) is an example of a creative application of the design with birthday as the IV. For encouragement designs, see Holland (1988) and Imai et al. (2011) .

MATCHING AND PROPENSITY SCORE DESIGN

This section considers quasi-experimental designs in which researchers lack control over treatment selection but have good knowledge about the selection mechanism or at least the confounders that simultaneously determine the treatment selection and the outcome. Due to self or third-person selection of subjects into treatment, the resulting treatment and control groups typically differ in observed but also unobserved baseline covariates. If we have reliable measures of all confounding covariates, then matching or propensity score (PS) designs balance groups on observed baseline covariates and thus enable the identification of causal effects ( Imbens & Rubin, 2015 ). Regression analysis and the analysis of covariance can also remove the confounding bias, but because they rely on functional form assumptions and extrapolation we discuss only nonparametric matching and PS designs.

Suppose that students decide on their own whether to attend the science camp. Although many factors can affect students’ decision, teachers with several years of experience of running the camp may know that selection is mostly driven by students’ science ability, liking of science, and their parents’ socioeconomic status. If all the selection-relevant factors that also affect the outcome are known, the question mark in Figure 2 can be replaced by the known confounding covariates.

Given the set of confounding covariates, causal inference with matching or PS designs is straightforward, at least theoretically. The basic one-to-one matching design matches each treatment subject to a control subject that is equivalent or at least very similar in observed covariates. To illustrate the idea of matching, consider a camp attendee with baseline measures of 80 on the science pre-test, 6 on liking science, and 50 on the socioeconomic status. Then a multivariate matching strategy tries to find a nonattendee with exactly the same or at least very similar baseline measures. If we succeed in finding close matches for all camp attendee, the matched samples of attendees and nonattendees will have almost identical covariate distributions.

Although multivariate matching works well when the number of confounders is small and the pool of control subjects is large relative to the number of treatment subjects, it is usually difficult to find close matches with a large set of covariates or a small pool of control subjects. Matching on the PS helps to overcome this issue because the PS is a univariate score computed from the observed covariates ( Rosenbaum & Rubin, 1983 ). The PS is formally defined as the conditional probability of receiving the treatment given the set of observed covariates X : PS = Pr( Z = 1 | X ).

Matching and PS designs usually investigate ATE = E [ Y i (1)] − E [ Y i (0)] or ATT = E [ Y i (1) | Z i = 1] – E [ Y i (0) | Z i = 1]. Both causal effects are identified if (a) the potential outcomes are statistically independent of the treatment indicator given the set of observed confounders X , { Y (1), Y (0)}⊥ Z | X ( unconfoundedness ; ⊥ denotes independence), and (b) the treatment probability is strictly between zero and one, 0 < Pr( Z = 1 | X ) < 1 ( positivity ).

By the positivity assumption we get E [ Y i (1)] = E X [ E [ Y i (1) | X ]] and E [ Y i (0)] = E X [ E [ Y i (0) | X ]]. If the unconfoundedness assumption holds, we can write the inner expectations as E [ Y i (1) | X ] = E [ Y i (1) | Z i =1; X ] and E [ Y i (0) | X ] = E [ Y i (0) | Z i = 0; X ]. Finally, because the treatment (control) outcomes of the treatment (control) subjects are actually observed, ATE is identified because it can be expressed in terms of observable quantities: ATE = E X [ E [ Y i | Z i = 1; X ]] – E X [ E [ Y i | Z i = 0; X ]]. The same can be shown for ATT. The unconfoundedness and positivity assumption are frequently referred to jointly as the strong ignorability assumption. Rosenbaum and Rubin (1983) proved that if the assignment is strongly ignorable given X , then it is also strongly ignorable given the PS alone.

Estimating ATE and ATT

Matching designs use a distance measure for matching each treatment subject to the closest control subject. The Mahalanobis distance is usually used for multivariate matching and the Euclidean distance on the logit of the PS for PS matching. Matching strategies differ with respect to the matching ratio (one-to-one or one-to-many), replacement of matched subjects (with or without replacement), use of a caliper (treatment subjects that do not have a control subject within a certain threshold remain unmatched), and the matching algorithm (greedy, genetic, or optimal matching; Sekhon, 2011 ; Steiner & Cook, 2013 ). Because we try to find at least one control subject for each treatment subject, matching estimators typically estimate ATT. Once treatment and control subjects are matched, ATT is computed as the difference in the mean outcome of the treatment and control group. An alternative matching strategy that allows for estimating ATE is full matching, which stratifies all subjects into the maximum number of strata, where each stratum contains at least one treatment and one control subject ( Hansen, 2004 ).

The PS can also be used for PS stratification and inverse-propensity weighting. PS stratification stratifies the treatment and control subjects into at least five strata and estimates the treatment effect within each stratum. ATE or ATT is then obtained as the weighted average of the stratum-specific treatment effects. Inverse-propensity weighting follows the same logic as inverse-probability weighting in survey research ( Horvitz & Thompson, 1952 ) and requires the computation of weights that refer to either the overall population (ATE) or the population of treated subjects only (ATT). Given the inverse-propensity weights, ATE or ATT is usually estimated via weighted least squares regression.

Because the true PSs are unknown, they need to be estimated from the observed data. The most common method for estimating the PS is logistic regression, which regresses the binary treatment indicator Z on predictors of the observed covariates. The PS model is specified according to balance criteria (instead of goodness of fit criteria), that is, the estimated PSs should remove all baseline differences in observed covariates ( Imbens & Rubin, 2015 ). The predicted probabilities from the PS model represent the estimated PSs.

All three PS designs—matching, stratification, and weighting—can benefit from additional covariance adjustments in an outcome regression. That is, for the matched, stratified or weighted data, the outcome is regressed on the treatment indicator and the additional covariates. Combining the PS design with a covariance adjustment gives researchers two chances to remove the confounding bias, by correctly specifying either the PS model or the outcome model. These combined methods are said to be doubly robust because they are robust against either the misspecification of the PS model or the misspecification of the outcome model ( Robins & Rotnitzky, 1995 ). The R packages optmatch ( Hansen & Klopfer, 2006 ) and MatchIt ( Ho et al., 2011 ) and the STATA command teffects , in particular teffects psmatch ( StataCorp, 2015 ), can be useful for matching or PS analyses.

The most challenging issue with matching and PS designs is the selection of covariates for establishing unconfoundedness. Ideally, subject-matter theory about the selection process and the outcome-generating model is used for selecting a set of covariates that removes all the confounding ( Pearl, 2009 ). If strong subject-matter theories are not available, selecting the right covariates is difficult. In the hope to remove a major part of the confounding bias—if not all of it—a frequently applied strategy is to match on as many covariates as possible. However, recent literature shows that thoughtless inclusion of covariates may increase rather than reduce the confounding bias ( Pearl, 2010 ; Steiner & Kim, in press). The risk of increasing bias can be reduced if the observed covariates cover a broad range of heterogeneous construct domains, including at least one reliable pretest measure of the outcome ( Steiner, Cook, et al., 2015 ). Besides having the right covariates, they also need to be reliably measured. The unreliable measurement of confounding covariates has a similar effect as the omission of a confounder: It results in a violation of the unconfoundedness assumption and thus in a biased effect estimate ( Steiner, Cook, & Shadish, 2011 ; Steiner & Kim, in press ).

Even if the set of reliably measured covariates establishes unconfoundedness, we still need to correctly specify the functional form of the PS model. Although parametric models like logistic regression, including higher order terms, might frequently approximate the correct functional form, they still rely on the linearity assumption. The linearity assumption can be relaxed if one estimates the PS with statistical learning algorithms like classification trees, neural networks, or the LASSO ( Keller, Kim, & Steiner, 2015 ; McCaffrey, Ridgeway, & Morral, 2004 ).

Strengthening Matching and PS Designs

The credibility of matching and PS designs heavily relies on the unconfoundedness assumption. Although empirically untestable, there are indirect ways for assessing unconfoundedness. First, unaffected (nonequivalent) outcomes that are known to be unaffected by the treatment can be used ( Shadish et al., 2002 ). For instance, we may expect that attendance in the science camp does not significantly affect the reading score. Thus, if we observe a significant group difference in the reading score after the PS adjustment, bias due to unobserved confounders (e.g., general intelligence) is still likely. Second, adding a second but conceptually different control group allows for a similar test as with the unaffected outcome ( Rosenbaum, 2002 ).

Because researchers rarely know whether the unconfoundedness assumption is actually met with the data at hand, it is important to assess the effect estimate’s sensitivity to potentially unobserved confounders. Sensitivity analyses investigate how strongly an estimate’s magnitude and significance changes if a confounder of a certain strength would have been omitted from the analyses. Causal conclusions are much more credible if the effect’s direction, magnitude, and significance is rather insensitive to omitted confounders ( Rosenbaum, 2002 ). However, despite the value of sensitivity analyses, they are not informative about whether hidden bias is actually present.

Schafer and Kang (2008) and Steiner and Cook (2013) provided a comprehensive introduction. Rigorous formalization and technical details of PS designs can be found in Imbens and Rubin (2015) . Rosenbaum (2002) discussed many important design issues in these designs.

COMPARATIVE INTERRUPTED TIME SERIES DESIGN

The designs discussed so far require researchers to have either full control over treatment assignment or reliable knowledge of the exogenous (IV) or endogenous part of the selection mechanism (i.e., the confounders). If none of these requirements are met, a comparative interrupted time series (CITS) design might be a viable alternative if (a) multiple measurements of the outcome ( time series ) are available for both the treatment and a comparison group and (b) the treatment group’s time series has been interrupted by an intervention.

Suppose that all students of one class in a school (say, an advanced science class) attend the camp, whereas all students of another class in the same school do not attend. Also assume that monthly measures of science achievement before and after the science camp are available. Figure 3 illustrates such a scenario where the x -axis represents time in Months and the y -axis the Science Score (aggregated at the class level). The filled symbols indicate the treatment group (science camp), open symbols the comparison group (no science camp). The science camp intervention divides both time series into a preintervention time series (circles) and a postintervention time series (squares). The changes in the levels and slopes of the pre- and postintervention regression lines represent the camp’s impact but possibly also the effect of other events that co-occur with the intervention. The dashed lines extrapolate the preintervention growth curves into the postintervention period, and thus represent the counterfactual situation where the intervention but also other co-occurring events are absent.

An external file that holds a picture, illustration, etc.
Object name is nihms-983980-f0003.jpg

A hypothetical example of comparative interrupted time series design.

The strength of a CITS design is its ability to discriminate between the intervention’s effect and the effects of co-occurring events. Such events might be other potentially competing interventions (history effects) or changes in the measurement of the outcome (instrumentation), for instance. If the co-occurring events affect the treatment and comparison group to the same extent, then subtracting the changes in the comparison group’s growth curve from the changes in the treatment group’s growth curve provides a valid estimate of the intervention’s impact. Because we investigate the difference in the changes (= differences) of the two growth curves, the CITS design is a special case of the difference-in-differences design ( Somers et al., 2013 ).

Assume that a daily TV series about Albert Einstein was broadcast in the evenings of the science camp week and that students of both classes were exposed to the same extent to the TV series. It follows that the comparison group’s change in the growth curve represents the TV series’ impact. The comparison group’s time series in Figure 3 indicates that the TV series might have had an immediate impact on the growth curve’s level but almost no effect on the slope. On the other hand, the treatment group’s change in the growth curve is due to both the science camp and the TV series. Thus, in differencing out the TV series’ effect (estimated from the comparison group) we can identify the camp effect.

Let t c denote the time point of the intervention, then the intervention’s effect on the treated (ATT) at a postintervention time point t ≥ t c is defined as τ t = E [ Y i t T ( 1 ) ] − E [ Y i t T ( 0 ) ] , where Y i t T ( 0 ) and Y i t T ( 1 ) are the potential control and treatment outcomes of subject i in the treatment group ( T ) at time point t . The time series of the expected potential outcomes can be formalized as sum of nonparametric but additive time-dependent functions. The treatment group’s expected potential control outcome can be represented as E [ Y i t T ( 0 ) ] = f 0 T ( t ) + f E T ( t ) , where the control function f 0 T ( t ) generates the expected potential control outcomes in absence of any interventions ( I ) or co-occurring events ( E ), and the event function f E T ( t ) adds the effects of co-occurring events. Similarly, the expected potential treatment outcome can be written as E [ Y i t T ( 1 ) ] = f 0 T ( t ) + f E T ( t ) + f I T ( t ) , which adds the intervention’s effect τ t = f I T ( t ) to the control and event function. In the absence of a comparison group, we can try to identify the impact of the intervention by comparing the observable postintervention outcomes to the extrapolated outcomes from the preintervention time series (dashed line in Figure 3 ). Extrapolation is necessary because we do not observe any potential control outcomes in the postintervention period (only potential treatment outcomes are observed). Let f ^ 0 T ( t ) denote the parametric extrapolation of the preintervention control function f 0 T ( t ) , then the observable pre–post-intervention difference ( PP T ) in the expected control outcome is P P t T = f 0 T ( t ) + f E T ( t ) + f I T ( t ) − f ^ 0 T ( t ) = f I T ( t ) + ( f 0 T ( t ) − f ^ 0 T ( t ) ) + f E T ( t ) . Thus, in the absence of a comparison group, ATT is identified (i.e., P P t T = f I T ( t ) = τ t ) only if the control function is correctly specified ( f 0 T ( t ) = f ^ 0 T ( t ) ) and if no co-occurring events are present ( f E T ( t ) = 0 ).

The comparison group in a CITS design allows us to relax both of these identifying assumptions. In order to see this, we first define the expected control outcomes of the comparison group ( C ) as a sum of two time-dependent functions as before: E [ Y i t C ( 0 ) ] = f 0 C ( t ) + f E C ( t ) . Then, in extrapolating the comparison group’s preintervention function into the postintervention period, f ^ 0 C ( t ) , we can compute the pre–post-intervention difference for the comparison group: P P t C = f 0 C ( t ) + f E C ( t ) − f ^ 0 C ( t ) = f E C ( t ) + ( f 0 C ( t ) − f ^ 0 C ( t ) ) If the control function is correctly specified f 0 C ( t ) = f ^ 0 C ( t ) , the effect of co-occurring events is identified P P t C = f E C ( t ) . However, we do not necessarily need a correctly specified control function, because in a CITS design we focus on the difference in the treatment and comparison group’s pre–post-intervention differences, that is, P P t T − P P t C = f I T ( t ) + { ( f 0 T ( t ) − f ^ 0 T ( t ) ) − ( f 0 C ( t ) − f ^ 0 C ( t ) ) } + { f E T ( t ) − f E C ( t ) } . Thus, ATT is identified, P P t T − P P t C = f I T ( t ) = τ t , if (a) both control functions are either correctly specified or misspecified to the same additive extent such that ( f 0 T ( t ) − f ^ 0 T ( t ) ) = ( f 0 C ( t ) − f ^ 0 C ( t ) ) ( no differential misspecification ) and (b) the effect of co-occurring events is identical in the treatment and comparison group, f E T ( t ) = f E C ( t ) ( no differential event effects ).

Estimating ATT

CITS designs are typically analyzed with linear regression models that regress the outcome Y on the centered time variable ( T – t c ), the intervention indicator Z ( Z = 0 if t < t c , otherwise Z = 1), the group indicator G ( G = 1 for the treatment group and G = 0 for the control group), and the corresponding two-way and three-way interactions:

Depending on the number of subjects in each group, fixed or random effects for the subjects are included as well (time fixed or random effect can also be considered). β ^ 5 estimates the intervention’s immediate effect at the onset of the intervention (change in intercept) and β ^ 7 the intervention’s effect on the growth rate (change in slope). The inclusion of dummy variables for each postintervention time point (plus their interaction with the intervention and group indicators) would allow for a direct estimation of the time-specific effects. If the time series are long enough (at least 100 time points), then a more careful modeling of the autocorrelation structure via time series models should be considered.

Compared to other designs, CITS designs heavily rely on extrapolation and thus on functional form assumptions. Therefore, it is crucial that the functional forms of the pre- and postintervention time series (including their extrapolations) are correctly specified or at least not differentially misspecified. With short time series or measurement points that inadequately capture periodical variations, the correct specification of the functional form is very challenging. Another specification aspect concerns serial dependencies among the data points. Failing to model serial dependencies can bias effect estimates and their standard errors such that significance tests might be misleading. Accounting for serial dependencies requires autoregressive models (e.g., ARIMA models), but the time series should have at least 100 time points ( West, Biesanz, & Pitts, 2000 ). Standard fixed effects or random effects models deal at least partially with the dependence structure. Robust standard errors (e.g., Huber-White corrected ones) or the bootstrap can also be used to account for dependency structures.

Events that co-occur with the intervention of interest, like history or instrumentation effects, are a major threat to the time series designs that lack a comparison group ( Shadish et al., 2002 ). CITS designs are rather robust to co-occurring events as long as the treatment and comparison groups are affected to the same additive extent. However, there is no guarantee that both groups are exposed to the same events and affected to the same extent. For example, if students who do not attend the camp are less likely to watch the TV series, its effect cannot be completely differenced out (unless the exposure to the TV series is measured). If one uses aggregated data like class or school averages of achievement scores, then differential compositional shifts over time can also invalidate the CITS design. Compositional shifts occur due to dropouts or incoming subjects over time.

Strengthening CITS Designs

If the treatment and comparison group’s preintervention time series are very different (different levels and slopes), then the assumption that history or instrumentation threats affect both groups to the same additive extent may not hold. Matching treatment and comparison subjects prior to the analysis can increase the plausibility of this assumption. Instead of using all nonparticipating students of the comparison class, we may select only those students who have a similar level and growth in the preintervention science scores as the students participating in the camp. We can also match on additional covariates like socioeconomic status or motivation levels. Multivariate or PS matching can be used for this purpose. If the two groups are similar, it is more likely that they are affected by co-occurring events to the same extent.

As with the matching and PS designs, using an unaffected outcome in CITS designs helps to probe the untestable assumptions ( Coryn & Hobson, 2011 ; Shadish et al., 2002 ). For instance, we might expect that attending the science camp does not affect students’ reading scores but that some validity threats (e.g., attrition) operate on both the reading and science outcome. If we find a significant camp effect on the reading score, the validity of the CITS design for evaluating the camp’s impact on the science score is in doubt.

Another strategy to avoid validity threats is to control the time point of the intervention if possible. Researchers can wait with the implementation of the treatment until they have enough preintervention measures for reliably estimating the functional form. They can also choose to intervene when threats to validity are less likely (avoiding the week of the TV series). Control over the intervention also allows researchers to introduce and remove the treatment in subsequent time intervals, maybe even with switching replications between two (or more) groups. If the treatment is effective, we expect that the pattern of the intervention scheme is directly reflected in the time series of the outcome (for more details, see Shadish et al., 2002 ; for the literature on single case designs, see Kazdin, 2011 ).

A comprehensive introduction to CITS design can be found in Shadish et al. (2002) , which also addresses many classical applications. For more technical details of its identification, refer to Lechner (2011) . Wong, Cook, and Steiner (2009) evaluated the effect of No Child Left Behind using a CITS design.

CONCLUDING REMARKS

This article discussed four of the strongest quasi-experimental designs for causal inference when randomized experiments are not feasible. For each design we highlighted the identification strategies and the required assumptions. In practice, it is crucial that the design assumptions are met, otherwise biased effect estimates result. Because most important assumptions like the exclusion restriction or the unconfoundedness assumption are not directly testable, researchers should always try to assess their plausibility via indirect tests and investigate the effect estimates’ sensitivity to violations of these assumptions.

Our discussion of RD, IV, PS, and CITS designs made it also very clear that, in comparison to RCTs, quasi-experimental designs rely on more or stronger assumptions. With prefect control over treatment assignment and treatment implementation (as in an RCT), causal inference is warranted by a minimal set of assumptions. But with limited control over and knowledge about treatment assignment and implementation, stronger assumptions are required and causal effects might be identifiable only for local subpopulations. Nonetheless, observational data sometimes meet the assumptions of a quasi-experimental design, at least approximately, such that causal conclusions are credible. If so, the estimates of quasi-experimental designs—which exploit naturally occurring selection processes and real-world implementations of the treatment—are frequently better generalizable than the results from a controlled laboratory experiment. Thus, if external validity is a major concern, the results of randomized experiments should always be complemented by findings from valid quasi-experiments.

  • Angrist JD, Imbens GW, & Rubin DB (1996). Identification of causal effects using instrumental variables . Journal of the American Statistical Association , 91 , 444–455. [ Google Scholar ]
  • Angrist JD, & Krueger AB (1992). The effect of age at school entry on educational attainment: An application of instrumental variables with moments from two samples . Journal of the American Statistical Association , 87 , 328–336. [ Google Scholar ]
  • Angrist JD, & Lavy V (1999). Using Maimonides’ rule to estimate the effect of class size on scholastic achievment . Quarterly Journal of Economics , 114 , 533–575. [ Google Scholar ]
  • Angrist JD, & Pischke JS (2009). Mostly harmless econometrics: An empiricist’s companion . Princeton, NJ: Princeton University Press. [ Google Scholar ]
  • Angrist JD, & Pischke JS (2015). Mastering’metrics: The path from cause to effect . Princeton, NJ: Princeton University Press. [ Google Scholar ]
  • Baum CF, Schaffer ME, & Stillman S (2007). Enhanced routines for instrumental variables/generalized method of moments estimation and testing . The Stata Journal , 7 , 465–506. [ Google Scholar ]
  • Black D, Galdo J, & Smith JA (2007). Evaluating the bias of the regression discontinuity design using experimental data (Working paper) . Chicago, IL: University of Chicago. [ Google Scholar ]
  • Brito C, & Pearl J (2002). Generalized instrumental variables In Darwiche A & Friedman N (Eds.), Uncertainty in artificial intelligence (pp. 85–93). San Francisco, CA: Morgan Kaufmann. [ Google Scholar ]
  • Calonico S, Cattaneo MD, & Titiunik R (2015). rdrobust: Robust data-driven statistical inference in regression-discontinuity designs (R package ver. 0.80) . Retrieved from http://CRAN.R-project.org/package=rdrobust
  • Coryn CLS, & Hobson KA (2011). Using nonequivalent dependent variables to reduce internal validity threats in quasi-experiments: Rationale, history, and examples from practice . New Directions for Evaluation , 131 , 31–39. [ Google Scholar ]
  • Dimmery D (2013). rdd: Regression discontinuity estimation (R package ver. 0.56) . Retrieved from http://CRAN.R-project.org/package=rdd
  • Ding P, & Miratrix LW (2015). To adjust or not to adjust? Sensitivity analysis of M-bias and butterfly-bias . Journal of Causal Inference , 3 ( 1 ), 41–57. [ Google Scholar ]
  • Fox J (2006). Structural equation modeling with the sem package in R . Structural Equation Modeling , 13 , 465–486. [ Google Scholar ]
  • Hahn J, Todd P, & Van der Klaauw W (2001). Identification and estimation of treatment effects with a regression–discontinuity design . Econometrica , 69 ( 1 ), 201–209. [ Google Scholar ]
  • Hansen BB (2004). Full matching in an observational study of coaching for the SAT . Journal of the American Statistical Association , 99 , 609–618. [ Google Scholar ]
  • Hansen BB, & Klopfer SO (2006). Optimal full matching and related designs via network flows . Journal of Computational and Graphical Statistics , 15 , 609–627. [ Google Scholar ]
  • Ho D, Imai K, King G, & Stuart EA (2011). MatchIt: Nonparametric preprocessing for parametric causal inference . Journal of Statistical Software , 42 ( 8 ), 1–28. Retrieved from http://www.jstatsoft.org/v42/i08/ [ Google Scholar ]
  • Holland PW (1986). Statistics and causal inference . Journal of the American Statistical Association , 81 , 945–960. [ Google Scholar ]
  • Holland PW (1988). Causal inference, path analysis and recursive structural equations models . ETS Research Report Series . doi: 10.1002/j.2330-8516.1988.tb00270.x [ CrossRef ] [ Google Scholar ]
  • Horvitz DG, & Thompson DJ (1952). A generalization of sampling without replacement from a finite universe . Journal of the American Statistical Association , 47 , 663–685. [ Google Scholar ]
  • Imai K, Keele L, Tingley D, & Yamamoto T (2011). Unpacking the black box of causality: Learning about causal mechanisms from experimental and observational studies . American Political Science Review , 105 , 765–789. [ Google Scholar ]
  • Imbens GW, & Lemieux T (2008). Regression discontinuity designs: A guide to practice . Journal of Econometrics , 142 , 615–635. [ Google Scholar ]
  • Imbens GW, & Rubin DB (2015). Causal inference in statistics, social, and biomedical sciences . New York, NY: Cambridge University Press. [ Google Scholar ]
  • Kazdin AE (2011). Single-case research designs: Methods for clinical and applied settings . New York, NY: Oxford University Press. [ Google Scholar ]
  • Keller B, Kim JS, & Steiner PM (2015). Neural networks for propensity score estimation: Simulation results and recommendations In van der Ark LA, Bolt DM, Chow S-M, Douglas JA, & Wang W-C (Eds.), Quantitative psychology research (pp. 279–291). New York, NY: Springer. [ Google Scholar ]
  • Lechner M (2011). The estimation of causal effects by difference-in-difference methods . Foundations and Trends in Econometrics , 4 , 165–224. [ Google Scholar ]
  • Lee DS, & Lemieux T (2010). Regression discontinuity designs in economics . Journal of Economic Literature , 48 , 281–355. [ Google Scholar ]
  • McCaffrey DF, Ridgeway G, & Morral AR (2004). Propensity score estimation with boosted regression for evaluating causal effects in observational studies . Psychological Methods , 9 , 403–425. [ PubMed ] [ Google Scholar ]
  • McCrary J (2008). Manipulation of the running variable in the regression discontinuity design: A density test . Journal of Econometrics , 142 , 698–714. [ Google Scholar ]
  • Nichols A (2007). rd: Stata modules for regression discontinuity estimation . Retrieved from http://ideas.repec.org/c/boc/bocode/s456888.html
  • Pearl J (2009). C ausality: Models, reasoning, and inference (2nd ed.). New York, NY: Cambridge University Press. [ Google Scholar ]
  • Pearl J (2010). On a class of bias-amplifying variables that endanger effect estimates In Proceedings of the Twenty-Sixth Conference on Uncertainty in Artificial Intelligence (pp. 425–432). Corvallis, OR: Association for Uncertainty in Artificial Intelligence. [ Google Scholar ]
  • Robins JM, & Rotnitzky A (1995). Semiparametric efficiency in multivariate regression models with missing data . Journal of the American Statistical Association , 90 ( 429 ), 122–129. [ Google Scholar ]
  • Rosenbaum PR (2002). Observational studies . New York, NY: Springer. [ Google Scholar ]
  • Rosenbaum PR, & Rubin DB (1983). The central role of the propensity score in observational studies for causal effects . Biometrika , 70 ( 1 ), 41–55. [ Google Scholar ]
  • Schafer JL, & Kang J (2008). Average causal effects from nonrandomized studies: A practical guide and simulated example . Psychological Methods , 13 , 279–313. [ PubMed ] [ Google Scholar ]
  • Sekhon JS (2011). Multivariate and propensity score matching software with automated balance optimization: The matching package for R . Journal of Statistical Software , 42 ( 7 ), 1–52. [ Google Scholar ]
  • Shadish WR, Cook TD, & Campbell DT (2002). Experimental and quasi-experimental designs for generalized causal inference . Boston, MA: Houghton-Mifflin. [ Google Scholar ]
  • Somers M, Zhu P, Jacob R, & Bloom H (2013). The validity and precision of the comparative interrupted time series design and the difference-in-difference design in educational evaluation (MDRC working paper in research methodology) . New York, NY: MDRC. [ Google Scholar ]
  • StataCorp. (2015). Stata treatment-effects reference manual: Potential outcomes/counterfactual outcomes . College Station, TX: Stata Press; Retrieved from http://www.stata.com/manuals14/te.pdf [ Google Scholar ]
  • Steiner PM, & Cook D (2013). Matching and propensity scores In Little T (Ed.), The Oxford handbook of quantitative methods in psychology (Vol. 1 , pp. 237–259). New York, NY: Oxford University Press. [ Google Scholar ]
  • Steiner PM, Cook TD, Li W, & Clark MH (2015). Bias reduction in quasi-experiments with little selection theory but many covariates . Journal of Research on Educational Effectiveness , 8 , 552–576. [ Google Scholar ]
  • Steiner PM, Cook TD, & Shadish WR (2011). On the importance of reliable covariate measurement in selection bias adjustments using propensity scores . Journal of Educational and Behavioral Statistics , 36 , 213–236. [ Google Scholar ]
  • Steiner PM, & Kim Y (in press). The mechanics of omitted variable bias: Bias amplification and cancellation of offsetting biases . Journal of Causal Inference . [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Steiner PM, Kim Y, Hall CE, & Su D (2015). Graphical models for quasi-experimental designs . Sociological Methods & Research. Advance online publication . doi: 10.1177/0049124115582272 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • West SG, Biesanz JC, & Pitts SC (2000). Causal inference and generalization in field settings: Experimental and quasi-experimental designs In Reis HT & Judd CM (Eds.), Handbook of research methods in social and personality psychology (pp. 40–84). New York, NY: Cambridge University Press. [ Google Scholar ]
  • Wing C, & Cook TD (2013). Strengthening the regression discontinuity design using additional design elements: A within-study comparison . Journal of Policy Analysis and Management , 32 , 853–877. [ Google Scholar ]
  • Wong M, Cook TD, & Steiner PM (2009). No Child Left Behind: An interim evaluation of its effects on learning using two interrupted time series each with its own non-equivalent comparison series (Working Paper No. WP-09–11) . Evanston, IL: Institute for Policy Research, Northwestern University. [ Google Scholar ]
  • Wong VC, Wing C, Steiner PM, Wong M, & Cook TD (2012). Research designs for program evaluation . Handbook of Psychology , 2 , 316–341. [ Google Scholar ]
  • Wooldridge J (2012). Introductory econometrics: A modern approach (5th ed.). Mason, OH: South-Western Cengage Learning. [ Google Scholar ]

Logo for Open Library Publishing Platform

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

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:

The Effects of Psychotherapy: An Evaluation

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.

Figure 7.3 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.

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
  • 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. ↵

Research Methods in Psychology 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

IMAGES

  1. Research Design

    quasi experimental definition in research

  2. PPT

    quasi experimental definition in research

  3. PPT

    quasi experimental definition in research

  4. 5 Quasi-Experimental Design Examples (2024)

    quasi experimental definition in research

  5. Quasi-experimental Research: What It Is, Types & Examples

    quasi experimental definition in research

  6. Quasi-Experimental Design

    quasi experimental definition in research

VIDEO

  1. AI for Growth: A Proposed Quasi-Experimental Research Study (Using ITS Design)

  2. Quantitative Research. Experimental and Quasi

  3. Quasi experimental research design|3rd yr bsc nursing #notes #nursing #research

  4. Types of Quasi Experimental Research Design

  5. EXPERIMENTAL DESIGNS: TRUE AND QUASI DESIGNS

  6. Quasi-Experimental

COMMENTS

  1. Quasi-Experimental Design

    True experimental design Quasi-experimental design; Assignment to treatment: The researcher randomly assigns subjects to control and treatment groups.: Some other, non-random method is used to assign subjects to groups. Control over treatment: The researcher usually designs the treatment.: The researcher often does not have control over the treatment, but instead studies pre-existing groups ...

  2. Quasi-experimental Research: What It Is, Types & Examples

    Quasi-experimental research designs are a type of research design that is similar to experimental designs but doesn't give full control over the independent variable (s) like true experimental designs do. In a quasi-experimental design, the researcher changes or watches an independent variable, but the participants are not put into groups at ...

  3. 7.3 Quasi-Experimental 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.

  4. Quasi-Experimental Research 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 ...

  5. 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 ...

  6. Quasi Experimental Design Overview & Examples

    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 experiments. In these studies, researchers can manipulate an independent variable — that is, they change one factor to see what effect it has.

  7. 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.

  8. Quasi-Experimental Design: Definition, Types, Examples

    In traditional experimental designs, randomization is a powerful tool for ensuring that groups are equivalent at the outset of a study. However, quasi-experimental design often involves non-randomization due to the nature of the research. This means that participants are not randomly assigned to treatment and control groups.

  9. 14

    Specifically, we describe four quasi-experimental designs - one-group pretest-posttest designs, non-equivalent group designs, regression discontinuity designs, and interrupted time-series designs - and their statistical analyses in detail. Both simple quasi-experimental designs and embellishments of these simple designs are presented.

  10. Quasi-Experimental Design

    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 ...

  11. Quasi-Experimental Research: Types, Examples & Application

    Quasi-experimental research is a way of finding out if there's a cause-and-effect relationship between variables when true experiments are not possible because of practical or ethical constraints. For example, you want to know if a new medicine is effective for migraines. Instead of giving the medication to some people and not to others, you ...

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

    In medical informatics, the quasi-experimental, sometimes called the pre-post intervention, design often is used to evaluate the benefits of specific interventions. The increasing capacity of health care institutions to collect routine clinical data has led to the growing use of quasi-experimental study designs in the field of medical ...

  13. How to Use and Interpret Quasi-Experimental Design

    A quasi-experimental study (also known as a non-randomized pre-post intervention) is a research design in which the independent variable is manipulated, but participants are not randomly assigned to conditions. Commonly used in medical informatics (a field that uses digital information to ensure better patient care), researchers generally use ...

  14. Quasi-Experimental Research

    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 ...

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

    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.

  16. Quasi-Experimental Design (SOCIAL PSYCHOLOGY)

    Quasi-Experimental Design Definition. A quasi-experimental design is a research methodology that possesses some, but not all, of the defining characteristics of a true experiment. In most cases, such designs examine the impact of one or more independent variables on dependent variables, but without assigning participants to conditions randomly ...

  17. Experimental and Quasi-Experimental Designs in Implementation Research

    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.

  18. Use of Quasi-Experimental Research Designs in Education Research

    In the past few decades, we have seen a rapid proliferation in the use of quasi-experimental research designs in education research. This trend, stemming in part from the "credibility revolution" in the social sciences, particularly economics, is notable along with the increasing use of randomized controlled trials in the strive toward rigorous causal inference.

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

    Quasi-experimental designs (QEDs) are increasingly employed to achieve a better balance between internal and external validity. Although these designs are often referred to and summarized in terms of logistical benefits versus threats to internal validity, there is still uncertainty about: (1) how to select from among various QEDs, and (2 ...

  20. Quasi-Experimental Designs for Causal Inference

    This article discusses four of the strongest quasi-experimental designs for identifying causal effects: regression discontinuity design, instrumental variable design, matching and propensity score designs, and the comparative interrupted time series design. For each design we outline the strategy and assumptions for identifying a causal effect ...

  21. Quasi-Experimental Research

    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 ...