User Preferences

Content preview.

Arcu felis bibendum ut tristique et egestas quis:

  • Ut enim ad minim veniam, quis nostrud exercitation ullamco laboris
  • Duis aute irure dolor in reprehenderit in voluptate
  • Excepteur sint occaecat cupidatat non proident

Keyboard Shortcuts

6.6 - confidence intervals & hypothesis testing.

Confidence intervals and hypothesis tests are similar in that they are both inferential methods that rely on an approximated sampling distribution. Confidence intervals use data from a sample to estimate a population parameter. Hypothesis tests use data from a sample to test a specified hypothesis. Hypothesis testing requires that we have a hypothesized parameter. 

The simulation methods used to construct bootstrap distributions and randomization distributions are similar. One primary difference is a bootstrap distribution is centered on the observed sample statistic while a randomization distribution is centered on the value in the null hypothesis. 

In Lesson 4, we learned confidence intervals contain a range of reasonable estimates of the population parameter. All of the confidence intervals we constructed in this course were two-tailed. These two-tailed confidence intervals go hand-in-hand with the two-tailed hypothesis tests we learned in Lesson 5. The conclusion drawn from a two-tailed confidence interval is usually the same as the conclusion drawn from a two-tailed hypothesis test. In other words, if the the 95% confidence interval contains the hypothesized parameter, then a hypothesis test at the 0.05 \(\alpha\) level will almost always fail to reject the null hypothesis. If the 95% confidence interval does not contain the hypothesize parameter, then a hypothesis test at the 0.05 \(\alpha\) level will almost always reject the null hypothesis.

Example: Mean Section  

This example uses the Body Temperature dataset built in to StatKey for constructing a  bootstrap confidence interval and conducting a randomization test . 

Let's start by constructing a 95% confidence interval using the percentile method in StatKey:

  

The 95% confidence interval for the mean body temperature in the population is [98.044, 98.474].

Now, what if we want to know if there is enough evidence that the mean body temperature is different from 98.6 degrees? We can conduct a hypothesis test. Because 98.6 is not contained within the 95% confidence interval, it is not a reasonable estimate of the population mean. We should expect to have a p value less than 0.05 and to reject the null hypothesis.

\(H_0: \mu=98.6\)

\(H_a: \mu \ne 98.6\)

\(p = 2*0.00080=0.00160\)

\(p \leq 0.05\), reject the null hypothesis

There is evidence that the population mean is different from 98.6 degrees. 

Selecting the Appropriate Procedure Section  

The decision of whether to use a confidence interval or a hypothesis test depends on the research question. If we want to estimate a population parameter, we use a confidence interval. If we are given a specific population parameter (i.e., hypothesized value), and want to determine the likelihood that a population with that parameter would produce a sample as different as our sample, we use a hypothesis test. Below are a few examples of selecting the appropriate procedure. 

Example: Cheese Consumption Section  

Research question: How much cheese (in pounds) does an average American adult consume annually? 

What is the appropriate inferential procedure? 

Cheese consumption, in pounds, is a quantitative variable. We have one group: American adults. We are not given a specific value to test, so the appropriate procedure here is a  confidence interval for a single mean .

Example: Age Section  

Research question:  Is the average age in the population of all STAT 200 students greater than 30 years?

There is one group: STAT 200 students. The variable of interest is age in years, which is quantitative. The research question includes a specific population parameter to test: 30 years. The appropriate procedure is a  hypothesis test for a single mean .

Try it! Section  

For each research question, identify the variables, the parameter of interest and decide on the the appropriate inferential procedure.

Research question:  How strong is the correlation between height (in inches) and weight (in pounds) in American teenagers?

There are two variables of interest: (1) height in inches and (2) weight in pounds. Both are quantitative variables. The parameter of interest is the correlation between these two variables.

We are not given a specific correlation to test. We are being asked to estimate the strength of the correlation. The appropriate procedure here is a  confidence interval for a correlation . 

Research question:  Are the majority of registered voters planning to vote in the next presidential election?

The parameter that is being tested here is a single proportion. We have one group: registered voters. "The majority" would be more than 50%, or p>0.50. This is a specific parameter that we are testing. The appropriate procedure here is a  hypothesis test for a single proportion .

Research question:  On average, are STAT 200 students younger than STAT 500 students?

We have two independent groups: STAT 200 students and STAT 500 students. We are comparing them in terms of average (i.e., mean) age.

If STAT 200 students are younger than STAT 500 students, that translates to \(\mu_{200}<\mu_{500}\) which is an alternative hypothesis. This could also be written as \(\mu_{200}-\mu_{500}<0\), where 0 is a specific population parameter that we are testing. 

The appropriate procedure here is a  hypothesis test for the difference in two means .

Research question:  On average, how much taller are adult male giraffes compared to adult female giraffes?

There are two groups: males and females. The response variable is height, which is quantitative. We are not given a specific parameter to test, instead we are asked to estimate "how much" taller males are than females. The appropriate procedure is a  confidence interval for the difference in two means .

Research question:  Are STAT 500 students more likely than STAT 200 students to be employed full-time?

There are two independent groups: STAT 500 students and STAT 200 students. The response variable is full-time employment status which is categorical with two levels: yes/no.

If STAT 500 students are more likely than STAT 200 students to be employed full-time, that translates to \(p_{500}>p_{200}\) which is an alternative hypothesis. This could also be written as \(p_{500}-p_{200}>0\), where 0 is a specific parameter that we are testing. The appropriate procedure is a  hypothesis test for the difference in two proportions.

Research question:  Is there is a relationship between outdoor temperature (in Fahrenheit) and coffee sales (in cups per day)?

There are two variables here: (1) temperature in Fahrenheit and (2) cups of coffee sold in a day. Both variables are quantitative. The parameter of interest is the correlation between these two variables.

If there is a relationship between the variables, that means that the correlation is different from zero. This is a specific parameter that we are testing. The appropriate procedure is a  hypothesis test for a correlation . 

Have a language expert improve your writing

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

  • Knowledge Base
  • Understanding Confidence Intervals | Easy Examples & Formulas

Understanding Confidence Intervals | Easy Examples & Formulas

Published on August 7, 2020 by Rebecca Bevans . Revised on June 22, 2023.

When you make an estimate in statistics, whether it is a summary statistic or a test statistic , there is always uncertainty around that estimate because the number is based on a sample of the population you are studying.

The confidence interval is the range of values that you expect your estimate to fall between a certain percentage of the time if you run your experiment again or re-sample the population in the same way.

The confidence level is the percentage of times you expect to reproduce an estimate between the upper and lower bounds of the confidence interval, and is set by the alpha value .

Table of contents

What exactly is a confidence interval, calculating a confidence interval: what you need to know, confidence interval for the mean of normally-distributed data, confidence interval for proportions, confidence interval for non-normally distributed data, reporting confidence intervals, caution when using confidence intervals, other interesting articles, frequently asked questions about confidence intervals.

A confidence interval is the mean of your estimate plus and minus the variation in that estimate. This is the range of values you expect your estimate to fall between if you redo your test, within a certain level of confidence.

Confidence , in statistics, is another way to describe probability. For example, if you construct a confidence interval with a 95% confidence level, you are confident that 95 out of 100 times the estimate will fall between the upper and lower values specified by the confidence interval.

Your desired confidence level is usually one minus the alpha (α) value you used in your statistical test :

Confidence level = 1 − a

So if you use an alpha value of p < 0.05 for statistical significance , then your confidence level would be 1 − 0.05 = 0.95, or 95%.

When do you use confidence intervals?

You can calculate confidence intervals for many kinds of statistical estimates, including:

  • Proportions
  • Population means
  • Differences between population means or proportions
  • Estimates of variation among groups

These are all point estimates, and don’t give any information about the variation around the number. Confidence intervals are useful for communicating the variation around a point estimate.

However, the British people surveyed had a wide variation in the number of hours watched, while the Americans all watched similar amounts.

Even though both groups have the same point estimate (average number of hours watched), the British estimate will have a wider confidence interval than the American estimate because there is more variation in the data.

Variation around an estimate

Here's why students love Scribbr's proofreading services

Discover proofreading & editing

Most statistical programs will include the confidence interval of the estimate when you run a statistical test.

If you want to calculate a confidence interval on your own, you need to know:

  • The point estimate you are constructing the confidence interval for
  • The critical values for the test statistic
  • The standard deviation of the sample
  • The sample size

Once you know each of these components, you can calculate the confidence interval for your estimate by plugging them into the confidence interval formula that corresponds to your data.

Point estimate

The point estimate of your confidence interval will be whatever statistical estimate you are making (e.g., population mean , the difference between population means, proportions, variation among groups).

Finding the critical value

Critical values tell you how many standard deviations away from the mean you need to go in order to reach the desired confidence level for your confidence interval.

There are three steps to find the critical value.

  • Choose your alpha (α) value.

The alpha value is the probability threshold for statistical significance . The most common alpha value is p = 0.05, but 0.1, 0.01, and even 0.001 are sometimes used. It’s best to look at the research papers published in your field to decide which alpha value to use.

  • Decide if you need a one-tailed interval or a two-tailed interval.

You will most likely use a two-tailed interval unless you are doing a one-tailed t test .

For a two-tailed interval, divide your alpha by two to get the alpha value for the upper and lower tails.

  • Look up the critical value that corresponds with the alpha value.

If your data follows a normal distribution , or if you have a large sample size ( n > 30) that is approximately normally distributed, you can use the z distribution to find your critical values.

For a z statistic, some of the most common values are shown in this table:

Confidence level 90% 95% 99%
alpha for one-tailed CI 0.1 0.05 0.01
alpha for two-tailed CI 0.05 0.025 0.005
statistic 1.64 1.96 2.57

If you are using a small dataset (n ≤ 30) that is approximately normally distributed, use the t distribution instead.

The t distribution follows the same shape as the z distribution, but corrects for small sample sizes. For the t distribution, you need to know your degrees of freedom (sample size minus 1).

Check out this set of t tables to find your t statistic. We have included the confidence level and p values for both one-tailed and two-tailed tests to help you find the t value you need.

For normal distributions, like the t distribution and z distribution, the critical value is the same on either side of the mean.

For a two-tailed 95% confidence interval, the alpha value is 0.025, and the corresponding critical value is 1.96.

Finding the standard deviation

Most statistical software will have a built-in function to calculate your standard deviation, but to find it by hand you can first find your sample variance, then take the square root to get the standard deviation.

  • Find the sample variance

Sample variance is defined as the sum of squared differences from the mean, also known as the mean-squared-error (MSE):

s^2 = {\sum}^n _{i=1} {\frac {(Xi - \bar{X})^2}{n-1}}

To find the MSE, subtract your sample mean from each value in the dataset, square the resulting number, and divide that number by n − 1 (sample size minus 1).

Then add up all of these numbers to get your total sample variance ( s 2 ). For larger sample sets, it’s easiest to do this in Excel.

  • Find the standard deviation.

The standard deviation of your estimate ( s ) is equal to the square root of the sample variance/sample error ( s 2 ):

s = \sqrt{(s^2)}

  • 10 for the GB estimate.
  • 5 for the USA estimate.

Sample size

The sample size is the number of observations in your data set.

Normally-distributed data forms a bell shape when plotted on a graph, with the sample mean in the middle and the rest of the data distributed fairly evenly on either side of the mean.

The confidence interval for data which follows a standard normal distribution is:

CI = \bar{X} \pm Z^* \frac {\sigma}{\sqrt{n}}

  • CI = the confidence interval
  • X̄ = the population mean
  • Z* = the critical value of the z distribution
  • σ = the population standard deviation
  • √n = the square root of the population size

The confidence interval for the t distribution follows the same formula, but replaces the Z * with the t *.

In real life, you never know the true values for the population (unless you can do a complete census). Instead, we replace the population values with the values from our sample data, so the formula becomes:

CI = \hat{x} \pm Z^* \frac {s}{\sqrt{n}}

  • ˆx = the sample mean
  • s = the sample standard deviation

To calculate the 95% confidence interval, we can simply plug the values into the formula.

For the USA:

\begin{align*} CI &= 35 \pm 1.96 \dfrac{5}{\sqrt{100}} \\ &= 35 \pm 1.96(0.5) \\ &= 35 \pm 0.98 \end{align*}

So for the USA, the lower and upper bounds of the 95% confidence interval are 34.02 and 35.98.

\begin{align*} CI &= 35 \pm 1.96 \dfrac{10}{\sqrt{100}} \\ &= 35 \pm 1.96(1) \\ &= 35 \pm 1.96 \end{align*}

The confidence interval for a proportion follows the same pattern as the confidence interval for means, but place of the standard deviation you use the sample proportion times one minus the proportion:

CI = \hat{p} \pm Z^* \sqrt{\dfrac{{\hat{p}(1-\hat{p})}}{n}}

  • ˆp = the proportion in your sample (e.g. the proportion of respondents who said they watched any television at all)
  • Z*= the critical value of the z distribution
  • n = the sample size

Prevent plagiarism. Run a free check.

To calculate a confidence interval around the mean of data that is not normally distributed, you have two choices:

  • You can find a distribution that matches the shape of your data and use that distribution to calculate the confidence interval.
  • You can perform a transformation on your data to make it fit a normal distribution, and then find the confidence interval for the transformed data.

Performing data transformations is very common in statistics, for example, when data follows a logarithmic curve but we want to use it alongside linear data. You just have to remember to do the reverse transformation on your data when you calculate the upper and lower bounds of the confidence interval.

Confidence intervals are sometimes reported in papers, though researchers more often report the standard deviation of their estimate.

If you are asked to report the confidence interval, you should include the upper and lower bounds of the confidence interval.

One place that confidence intervals are frequently used is in graphs. When showing the differences between groups, or plotting a linear regression, researchers will often include the confidence interval to give a visual representation of the variation around the estimate.

Confidence interval in a graph

Confidence intervals are sometimes interpreted as saying that the ‘true value’ of your estimate lies within the bounds of the confidence interval.

This is not the case. The confidence interval cannot tell you how likely it is that you found the true value of your statistical estimate because it is based on a sample, not on the whole population .

The confidence interval only tells you what range of values you can expect to find if you re-do your sampling or run your experiment again in the exact same way.

The more accurate your sampling plan, or the more realistic your experiment, the greater the chance that your confidence interval includes the true value of your estimate. But this accuracy is determined by your research methods, not by the statistics you do after you have collected the data!

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
  • Descriptive statistics
  • Measures of central tendency
  • Correlation coefficient

Methodology

  • Cluster sampling
  • Stratified sampling
  • Types of interviews
  • Cohort study
  • Thematic analysis

Research bias

  • Implicit bias
  • Cognitive bias
  • Survivorship bias
  • Availability heuristic
  • Nonresponse bias
  • Regression to the mean

The confidence level is the percentage of times you expect to get close to the same estimate if you run your experiment again or resample the population in the same way.

The confidence interval consists of the upper and lower bounds of the estimate you expect to find at a given level of confidence.

For example, if you are estimating a 95% confidence interval around the mean proportion of female babies born every year based on a random sample of babies, you might find an upper bound of 0.56 and a lower bound of 0.48. These are the upper and lower bounds of the confidence interval. The confidence level is 95%.

To calculate the confidence interval , you need to know:

Then you can plug these components into the confidence interval formula that corresponds to your data. The formula depends on the type of estimate (e.g. a mean or a proportion) and on the distribution of your data.

The standard normal distribution , also called the z -distribution, is a special normal distribution where the mean is 0 and the standard deviation is 1.

Any normal distribution can be converted into the standard normal distribution by turning the individual values into z -scores. In a z -distribution, z -scores tell you how many standard deviations away from the mean each value lies.

The z -score and t -score (aka z -value and t -value) show how many standard deviations away from the mean of the distribution you are, assuming your data follow a z -distribution or a t -distribution .

These scores are used in statistical tests to show how far from the mean of the predicted distribution your statistical estimate is. If your test produces a z -score of 2.5, this means that your estimate is 2.5 standard deviations from the predicted mean.

The predicted mean and distribution of your estimate are generated by the null hypothesis of the statistical test you are using. The more standard deviations away from the predicted mean your estimate is, the less likely it is that the estimate could have occurred under the null hypothesis .

A critical value is the value of the test statistic which defines the upper and lower bounds of a confidence interval , or which defines the threshold of statistical significance in a statistical test. It describes how far from the mean of the distribution you have to go to cover a certain amount of the total variation in the data (i.e. 90%, 95%, 99%).

If you are constructing a 95% confidence interval and are using a threshold of statistical significance of p = 0.05, then your critical value will be identical in both cases.

If your confidence interval for a difference between groups includes zero, that means that if you run your experiment again you have a good chance of finding no difference between groups.

If your confidence interval for a correlation or regression includes zero, that means that if you run your experiment again there is a good chance of finding no correlation in your data.

In both of these cases, you will also find a high p -value when you run your statistical test, meaning that your results could have occurred under the null hypothesis of no relationship between variables or no difference between groups.

If you want to calculate a confidence interval around the mean of data that is not normally distributed , you have two choices:

  • Find a distribution that matches the shape of your data and use that distribution to calculate the confidence interval.
  • Perform a transformation on your data to make it fit a normal distribution, and then find the confidence interval for the transformed data.

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.

Bevans, R. (2023, June 22). Understanding Confidence Intervals | Easy Examples & Formulas. Scribbr. Retrieved September 9, 2024, from https://www.scribbr.com/statistics/confidence-interval/

Is this article helpful?

Rebecca Bevans

Rebecca Bevans

Other students also liked, understanding p values | definition and examples, test statistics | definition, interpretation, and examples, how to calculate standard deviation (guide) | calculator & examples, what is your plagiarism score.

Logo for MacEwan Open Books

8.6 Relationship Between Confidence Intervals and Hypothesis Tests

Confidence intervals (CI) and hypothesis tests should give consistent results: we should not reject [latex]H_0[/latex] at the significance level [latex]\alpha[/latex] if the corresponding [latex](1 - \alpha) \times 100\%[/latex] confidence interval contains the hypothesized value [latex]\mu_0[/latex]. Two-sided confidence intervals correspond to two-tailed tests, upper-tailed confidence intervals correspond to right-tailed tests, and lower-tailed confidence intervals correspond to left-tailed tests.

A [latex](1 - \alpha) \times 100\%[/latex] two-sided [latex]t[/latex] confidence interval is given in the form [latex](\bar{x} - t_{\alpha / 2} \frac{s}{\sqrt{n}}, \bar{x} + t_{\alpha / 2} \frac{s}{\sqrt{n}})[/latex]. A [latex](1 - \alpha) \times 100\%[/latex] upper-tailed t confidence interval is given by [latex](\bar{x} - t_{\alpha} \frac{s}{\sqrt{n}}, \infty)[/latex] and the number [latex]\bar{x} - t_{\alpha} \frac{s}{\sqrt{n}}[/latex] is called the lower bound of the interval. A [latex](1 - \alpha) \times 100\%[/latex] lower-tailed t confidence interval is given by [latex](- \infty, \bar{x} + t_{\alpha} \frac{s}{\sqrt{n}})[/latex] and the number [latex]\bar{x} + t_{\alpha} \frac{s}{\sqrt{n}}[/latex] is called the upper bound of the interval. We can also use confidence intervals to make conclusions about hypothesis tests: reject the null hypothesis [latex]H_0[/latex] at the significance level [latex]\alpha[/latex] if the corresponding [latex](1 - \alpha) \times 100\%[/latex] confidence interval does not contain the hypothesized value [latex]\mu_0[/latex]. The relationship is summarized in the following table.

Table 8.3 : Relationship Between Confidence Interval and Hypothesis Test

Null hypothesis [latex]H_0: \mu = \mu_0[/latex] [latex]H_0: \mu \leq \mu_0[/latex] [latex]H_0: \mu \geq \mu_0[/latex]
Alternative [latex]H_a: \mu \neq \mu_0[/latex] [latex]H_a: \mu \: \gt \: \mu_0[/latex] [latex]H_a: \mu
[latex](1 - \alpha) \times 100\%[/latex] CI [latex](\bar{x} - t_{\alpha / 2} \frac{s}{\sqrt{n}}, \bar{x} + t_{\alpha / 2} \frac{s}{\sqrt{n}})[/latex] [latex](\bar{x} - t_{\alpha} \frac{s}{\sqrt{n}}, \infty)[/latex] [latex](- \infty, \bar{x} + t_{\alpha} \frac{s}{\sqrt{n}})[/latex]
Decision

null hypothesis confidence interval

Here is the reason we should reject [latex]H_0[/latex] if [latex]\mu_0[/latex] is outside the corresponding confidence interval.

Take the right-tailed test for example, we should reject [latex]H_0[/latex] if the observed test statistic [latex]t_o[/latex] falls in the rejection region, that is if [latex]t_o \geq t_{\alpha}[/latex]. This implies [latex]t_o = \frac{\bar{x} - \mu_0}{s / \sqrt{n}} \geq t_{\alpha} \Longrightarrow \mu_0 \leq \bar{x} - t_{\alpha} \frac{s}{\sqrt{n}}.[/latex] Given that the upper-tailed confidence interval for a right-tailed test is [latex](\bar{x} - t_{\alpha / 2} \frac{s}{\sqrt{n}}, \infty)[/latex], [latex]\mu_0 \leq \bar{x} - t_{\alpha} \frac{s}{\sqrt{n}}[/latex] means the value of [latex]\mu_0[/latex] is outside the confidence interval. The same rationale applies to two-tailed and left-tailed tests. Therefore, we can reject [latex]H_0[/latex] at the significance level [latex]\alpha[/latex] if [latex]\mu_0[/latex] is outside the corresponding (1– [latex]\alpha[/latex] )×100% confidence interval.

Example: Relationship Between Confidence Intervals and Hypothesis Tests

The ankle-brachial index (ABI) compares the blood pressure of a patient’s arm to the blood pressure of the patient’s leg. The ABI can be an indicator of different diseases, including arterial diseases. A healthy (or normal) ABI is 0.9 or greater. Researchers obtained the ABI of 100 women with peripheral arterial disease and obtained a mean ABI of 0.64 with a standard deviation of 0.15.

  • Set up the hypotheses: [latex]H_0: \mu \geq 0.9[/latex] versus [latex]H_a: \mu < 0.9[/latex].
  • The significance level is [latex]\alpha = 0.05[/latex].
  • Compute the value of the test statistic: [latex]t_o = \frac{\bar{x} - \mu_0}{s / \sqrt{n}} = \frac{0.64 - 0.9}{0.15 / \sqrt{100}} = \frac{-0.26}{0.015} = -17.333[/latex] with [latex]df = n-1 = 100 -1 = 99[/latex] (not given in Table IV, use 95, the closest one smaller than 99).
  • Find the P-value. For a left-tailed test, the P-value is the area to the left of the observed test statistic [latex]t_o[/latex]. [latex]\mbox{P-value} = P(t \leq t_o) = P(t \leq -17.333) = P(t \geq 17.333) 2.629(t_{0.005})[/latex].
  • Decision: Since the P- value [latex]< 0.005 < 0.05(\alpha)[/latex], we should reject the null hypothesis [latex]H_0[/latex].
  • Conclusion: At the 5% significance level, the data provide sufficient evidence that, on average, women with peripheral arterial disease have an unhealthy ABI.

[latex]\left( - \infty, \bar{x} + t_{\alpha} \frac{s}{\sqrt{n}} \right)= \left( - \infty, 0.64 + 1.661 \times \frac{0.15}{\sqrt{100}} \right) = (- \infty , 0.665)[/latex].

  • Does the interval in part b) support the conclusion in part a)? In part a), we reject [latex]H_0[/latex] and claim that the mean ABI is below 0.9 for women with peripheral arterial disease. In part b), we are 95% confident that the mean ABI is less than 0.9 since the entire confidence interval is below 0.9. In other words, the hypothesized value 0.9 is outside the corresponding confidence interval, we should reject the null. Therefore, the results obtained in parts a) and b) are consistent.

Introduction to Applied Statistics Copyright © 2024 by Wanhua Su is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License , except where otherwise noted.

Icon Partners

  • Quality Improvement
  • Talk To Minitab

Understanding Hypothesis Tests: Confidence Intervals and Confidence Levels

Topics: Hypothesis Testing , Data Analysis , Statistics

In this series of posts, I show how hypothesis tests and confidence intervals work by focusing on concepts and graphs rather than equations and numbers.  

Previously, I used graphs to show what statistical significance really means . In this post, I’ll explain both confidence intervals and confidence levels, and how they’re closely related to P values and significance levels.

How to Correctly Interpret Confidence Intervals and Confidence Levels

A confidence interval is a range of values that is likely to contain an unknown population parameter. If you draw a random sample many times, a certain percentage of the confidence intervals will contain the population mean. This percentage is the confidence level.

Most frequently, you’ll use confidence intervals to bound the mean or standard deviation, but you can also obtain them for regression coefficients, proportions, rates of occurrence (Poisson), and for the differences between populations.

Just as there is a common misconception of how to interpret P values , there’s a common misconception of how to interpret confidence intervals. In this case, the confidence level is not the probability that a specific confidence interval contains the population parameter.

The confidence level represents the theoretical ability of the analysis to produce accurate intervals if you are able to assess many intervals and you know the value of the population parameter. For a specific confidence interval from one study, the interval either contains the population value or it does not—there’s no room for probabilities other than 0 or 1. And you can't choose between these two possibilities because you don’t know the value of the population parameter.

"The parameter is an unknown constant and no probability statement concerning its value may be made."  —Jerzy Neyman, original developer of confidence intervals.

This will be easier to understand after we discuss the graph below . . .

With this in mind, how do you interpret confidence intervals?

Confidence intervals serve as good estimates of the population parameter because the procedure tends to produce intervals that contain the parameter. Confidence intervals are comprised of the point estimate (the most likely value) and a margin of error around that point estimate. The margin of error indicates the amount of uncertainty that surrounds the sample estimate of the population parameter.

In this vein, you can use confidence intervals to assess the precision of the sample estimate. For a specific variable, a narrower confidence interval [90 110] suggests a more precise estimate of the population parameter than a wider confidence interval [50 150].

Confidence Intervals and the Margin of Error

Let’s move on to see how confidence intervals account for that margin of error. To do this, we’ll use the same tools that we’ve been using to understand hypothesis tests. I’ll create a sampling distribution using probability distribution plots , the t-distribution , and the variability in our data. We'll base our confidence interval on the energy cost data set that we've been using.

When we looked at significance levels , the graphs displayed a sampling distribution centered on the null hypothesis value, and the outer 5% of the distribution was shaded. For confidence intervals, we need to shift the sampling distribution so that it is centered on the sample mean and shade the middle 95%.

Probability distribution plot that illustrates how a confidence interval works

The shaded area shows the range of sample means that you’d obtain 95% of the time using our sample mean as the point estimate of the population mean. This range [267 394] is our 95% confidence interval.

Using the graph, it’s easier to understand how a specific confidence interval represents the margin of error, or the amount of uncertainty, around the point estimate. The sample mean is the most likely value for the population mean given the information that we have. However, the graph shows it would not be unusual at all for other random samples drawn from the same population to obtain different sample means within the shaded area. These other likely sample means all suggest different values for the population mean. Hence, the interval represents the inherent uncertainty that comes with using sample data.

You can use these graphs to calculate probabilities for specific values. However, notice that you can’t place the population mean on the graph because that value is unknown. Consequently, you can’t calculate probabilities for the population mean, just as Neyman said!

Why P Values and Confidence Intervals Always Agree About Statistical Significance

You can use either P values or confidence intervals to determine whether your results are statistically significant. If a hypothesis test produces both, these results will agree.

The confidence level is equivalent to 1 – the alpha level. So, if your significance level is 0.05, the corresponding confidence level is 95%.

  • If the P value is less than your significance (alpha) level, the hypothesis test is statistically significant.
  • If the confidence interval does not contain the null hypothesis value, the results are statistically significant.
  • If the P value is less than alpha, the confidence interval will not contain the null hypothesis value.

For our example, the P value (0.031) is less than the significance level (0.05), which indicates that our results are statistically significant. Similarly, our 95% confidence interval [267 394] does not include the null hypothesis mean of 260 and we draw the same conclusion.

To understand why the results always agree, let’s recall how both the significance level and confidence level work.

  • The significance level defines the distance the sample mean must be from the null hypothesis to be considered statistically significant.
  • The confidence level defines the distance for how close the confidence limits are to sample mean.

Both the significance level and the confidence level define a distance from a limit to a mean. Guess what? The distances in both cases are exactly the same!

The distance equals the critical t-value * standard error of the mean . For our energy cost example data, the distance works out to be $63.57.

Imagine this discussion between the null hypothesis mean and the sample mean:

Null hypothesis mean, hypothesis test representative : Hey buddy! I’ve found that you’re statistically significant because you’re more than $63.57 away from me!

Sample mean, confidence interval representative : Actually, I’m significant because you’re more than $63.57 away from me !

Very agreeable aren’t they? And, they always will agree as long as you compare the correct pairs of P values and confidence intervals. If you compare the incorrect pair, you can get conflicting results, as shown by common mistake #1 in this post .

Closing Thoughts

In statistical analyses, there tends to be a greater focus on P values and simply detecting a significant effect or difference. However, a statistically significant effect is not necessarily meaningful in the real world. For instance, the effect might be too small to be of any practical value.

It’s important to pay attention to the both the magnitude and the precision of the estimated effect. That’s why I'm rather fond of confidence intervals. They allow you to assess these important characteristics along with the statistical significance. You'd like to see a narrow confidence interval where the entire range represents an effect that is meaningful in the real world.

If you like this post, you might want to read the previous posts in this series that use the same graphical framework:

  • Part One: Why We Need to Use Hypothesis Tests
  • Part Two: Significance Levels (alpha) and P values

For more about confidence intervals, read my post where I compare them to tolerance intervals and prediction intervals .

If you'd like to see how I made the probability distribution plot, please read: How to Create a Graphical Version of the 1-sample t-Test .

minitab-on-twitter

You Might Also Like

  • Trust Center

© 2023 Minitab, LLC. All Rights Reserved.

  • Terms of Use
  • Privacy Policy
  • Cookies Settings

U.S. flag

An official website of the United States government

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

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

  • Publications
  • Account settings

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

  • Advanced Search
  • Journal List

Logo of f1000res

  • PMC5635437.1 ; 2015 Aug 25
  • PMC5635437.2 ; 2016 Jul 13
  • ➤ PMC5635437.3; 2016 Oct 10

Null hypothesis significance testing: a short tutorial

Cyril pernet.

1 Centre for Clinical Brain Sciences (CCBS), Neuroimaging Sciences, The University of Edinburgh, Edinburgh, UK

Version Changes

Revised. amendments from version 2.

This v3 includes minor changes that reflect the 3rd reviewers' comments - in particular the theoretical vs. practical difference between Fisher and Neyman-Pearson. Additional information and reference is also included regarding the interpretation of p-value for low powered studies.

Peer Review Summary

Review dateReviewer name(s)Version reviewedReview status
Dorothy Vera Margaret Bishop Approved with Reservations
Stephen J. Senn Approved
Stephen J. Senn Approved with Reservations
Marcel ALM van Assen Not Approved
Daniel Lakens Not Approved

Although thoroughly criticized, null hypothesis significance testing (NHST) remains the statistical method of choice used to provide evidence for an effect, in biological, biomedical and social sciences. In this short tutorial, I first summarize the concepts behind the method, distinguishing test of significance (Fisher) and test of acceptance (Newman-Pearson) and point to common interpretation errors regarding the p-value. I then present the related concepts of confidence intervals and again point to common interpretation errors. Finally, I discuss what should be reported in which context. The goal is to clarify concepts to avoid interpretation errors and propose reporting practices.

The Null Hypothesis Significance Testing framework

NHST is a method of statistical inference by which an experimental factor is tested against a hypothesis of no effect or no relationship based on a given observation. The method is a combination of the concepts of significance testing developed by Fisher in 1925 and of acceptance based on critical rejection regions developed by Neyman & Pearson in 1928 . In the following I am first presenting each approach, highlighting the key differences and common misconceptions that result from their combination into the NHST framework (for a more mathematical comparison, along with the Bayesian method, see Christensen, 2005 ). I next present the related concept of confidence intervals. I finish by discussing practical aspects in using NHST and reporting practice.

Fisher, significance testing, and the p-value

The method developed by ( Fisher, 1934 ; Fisher, 1955 ; Fisher, 1959 ) allows to compute the probability of observing a result at least as extreme as a test statistic (e.g. t value), assuming the null hypothesis of no effect is true. This probability or p-value reflects (1) the conditional probability of achieving the observed outcome or larger: p(Obs≥t|H0), and (2) is therefore a cumulative probability rather than a point estimate. It is equal to the area under the null probability distribution curve from the observed test statistic to the tail of the null distribution ( Turkheimer et al. , 2004 ). The approach proposed is of ‘proof by contradiction’ ( Christensen, 2005 ), we pose the null model and test if data conform to it.

In practice, it is recommended to set a level of significance (a theoretical p-value) that acts as a reference point to identify significant results, that is to identify results that differ from the null-hypothesis of no effect. Fisher recommended using p=0.05 to judge whether an effect is significant or not as it is roughly two standard deviations away from the mean for the normal distribution ( Fisher, 1934 page 45: ‘The value for which p=.05, or 1 in 20, is 1.96 or nearly 2; it is convenient to take this point as a limit in judging whether a deviation is to be considered significant or not’). A key aspect of Fishers’ theory is that only the null-hypothesis is tested, and therefore p-values are meant to be used in a graded manner to decide whether the evidence is worth additional investigation and/or replication ( Fisher, 1971 page 13: ‘it is open to the experimenter to be more or less exacting in respect of the smallness of the probability he would require […]’ and ‘no isolated experiment, however significant in itself, can suffice for the experimental demonstration of any natural phenomenon’). How small the level of significance is, is thus left to researchers.

What is not a p-value? Common mistakes

The p-value is not an indication of the strength or magnitude of an effect . Any interpretation of the p-value in relation to the effect under study (strength, reliability, probability) is wrong, since p-values are conditioned on H0. In addition, while p-values are randomly distributed (if all the assumptions of the test are met) when there is no effect, their distribution depends of both the population effect size and the number of participants, making impossible to infer strength of effect from them.

Similarly, 1-p is not the probability to replicate an effect . Often, a small value of p is considered to mean a strong likelihood of getting the same results on another try, but again this cannot be obtained because the p-value is not informative on the effect itself ( Miller, 2009 ). Because the p-value depends on the number of subjects, it can only be used in high powered studies to interpret results. In low powered studies (typically small number of subjects), the p-value has a large variance across repeated samples, making it unreliable to estimate replication ( Halsey et al. , 2015 ).

A (small) p-value is not an indication favouring a given hypothesis . Because a low p-value only indicates a misfit of the null hypothesis to the data, it cannot be taken as evidence in favour of a specific alternative hypothesis more than any other possible alternatives such as measurement error and selection bias ( Gelman, 2013 ). Some authors have even argued that the more (a priori) implausible the alternative hypothesis, the greater the chance that a finding is a false alarm ( Krzywinski & Altman, 2013 ; Nuzzo, 2014 ).

The p-value is not the probability of the null hypothesis p(H0), of being true, ( Krzywinski & Altman, 2013 ). This common misconception arises from a confusion between the probability of an observation given the null p(Obs≥t|H0) and the probability of the null given an observation p(H0|Obs≥t) that is then taken as an indication for p(H0) (see Nickerson, 2000 ).

Neyman-Pearson, hypothesis testing, and the α-value

Neyman & Pearson (1933) proposed a framework of statistical inference for applied decision making and quality control. In such framework, two hypotheses are proposed: the null hypothesis of no effect and the alternative hypothesis of an effect, along with a control of the long run probabilities of making errors. The first key concept in this approach, is the establishment of an alternative hypothesis along with an a priori effect size. This differs markedly from Fisher who proposed a general approach for scientific inference conditioned on the null hypothesis only. The second key concept is the control of error rates . Neyman & Pearson (1928) introduced the notion of critical intervals, therefore dichotomizing the space of possible observations into correct vs. incorrect zones. This dichotomization allows distinguishing correct results (rejecting H0 when there is an effect and not rejecting H0 when there is no effect) from errors (rejecting H0 when there is no effect, the type I error, and not rejecting H0 when there is an effect, the type II error). In this context, alpha is the probability of committing a Type I error in the long run. Alternatively, Beta is the probability of committing a Type II error in the long run.

The (theoretical) difference in terms of hypothesis testing between Fisher and Neyman-Pearson is illustrated on Figure 1 . In the 1 st case, we choose a level of significance for observed data of 5%, and compute the p-value. If the p-value is below the level of significance, it is used to reject H0. In the 2 nd case, we set a critical interval based on the a priori effect size and error rates. If an observed statistic value is below and above the critical values (the bounds of the confidence region), it is deemed significantly different from H0. In the NHST framework, the level of significance is (in practice) assimilated to the alpha level, which appears as a simple decision rule: if the p-value is less or equal to alpha, the null is rejected. It is however a common mistake to assimilate these two concepts. The level of significance set for a given sample is not the same as the frequency of acceptance alpha found on repeated sampling because alpha (a point estimate) is meant to reflect the long run probability whilst the p-value (a cumulative estimate) reflects the current probability ( Fisher, 1955 ; Hubbard & Bayarri, 2003 ).

An external file that holds a picture, illustration, etc.
Object name is f1000research-4-10487-g0000.jpg

The figure was prepared with G-power for a one-sided one-sample t-test, with a sample size of 32 subjects, an effect size of 0.45, and error rates alpha=0.049 and beta=0.80. In Fisher’s procedure, only the nil-hypothesis is posed, and the observed p-value is compared to an a priori level of significance. If the observed p-value is below this level (here p=0.05), one rejects H0. In Neyman-Pearson’s procedure, the null and alternative hypotheses are specified along with an a priori level of acceptance. If the observed statistical value is outside the critical region (here [-∞ +1.69]), one rejects H0.

Acceptance or rejection of H0?

The acceptance level α can also be viewed as the maximum probability that a test statistic falls into the rejection region when the null hypothesis is true ( Johnson, 2013 ). Therefore, one can only reject the null hypothesis if the test statistics falls into the critical region(s), or fail to reject this hypothesis. In the latter case, all we can say is that no significant effect was observed, but one cannot conclude that the null hypothesis is true. This is another common mistake in using NHST: there is a profound difference between accepting the null hypothesis and simply failing to reject it ( Killeen, 2005 ). By failing to reject, we simply continue to assume that H0 is true, which implies that one cannot argue against a theory from a non-significant result (absence of evidence is not evidence of absence). To accept the null hypothesis, tests of equivalence ( Walker & Nowacki, 2011 ) or Bayesian approaches ( Dienes, 2014 ; Kruschke, 2011 ) must be used.

Confidence intervals

Confidence intervals (CI) are builds that fail to cover the true value at a rate of alpha, the Type I error rate ( Morey & Rouder, 2011 ) and therefore indicate if observed values can be rejected by a (two tailed) test with a given alpha. CI have been advocated as alternatives to p-values because (i) they allow judging the statistical significance and (ii) provide estimates of effect size. Assuming the CI (a)symmetry and width are correct (but see Wilcox, 2012 ), they also give some indication about the likelihood that a similar value can be observed in future studies. For future studies of the same sample size, 95% CI give about 83% chance of replication success ( Cumming & Maillardet, 2006 ). If sample sizes however differ between studies, CI do not however warranty any a priori coverage.

Although CI provide more information, they are not less subject to interpretation errors (see Savalei & Dunn, 2015 for a review). The most common mistake is to interpret CI as the probability that a parameter (e.g. the population mean) will fall in that interval X% of the time. The correct interpretation is that, for repeated measurements with the same sample sizes, taken from the same population, X% of times the CI obtained will contain the true parameter value ( Tan & Tan, 2010 ). The alpha value has the same interpretation as testing against H0, i.e. we accept that 1-alpha CI are wrong in alpha percent of the times in the long run. This implies that CI do not allow to make strong statements about the parameter of interest (e.g. the mean difference) or about H1 ( Hoekstra et al. , 2014 ). To make a statement about the probability of a parameter of interest (e.g. the probability of the mean), Bayesian intervals must be used.

The (correct) use of NHST

NHST has always been criticized, and yet is still used every day in scientific reports ( Nickerson, 2000 ). One question to ask oneself is what is the goal of a scientific experiment at hand? If the goal is to establish a discrepancy with the null hypothesis and/or establish a pattern of order, because both requires ruling out equivalence, then NHST is a good tool ( Frick, 1996 ; Walker & Nowacki, 2011 ). If the goal is to test the presence of an effect and/or establish some quantitative values related to an effect, then NHST is not the method of choice since testing is conditioned on H0.

While a Bayesian analysis is suited to estimate that the probability that a hypothesis is correct, like NHST, it does not prove a theory on itself, but adds its plausibility ( Lindley, 2000 ). No matter what testing procedure is used and how strong results are, ( Fisher, 1959 p13) reminds us that ‘ […] no isolated experiment, however significant in itself, can suffice for the experimental demonstration of any natural phenomenon'. Similarly, the recent statement of the American Statistical Association ( Wasserstein & Lazar, 2016 ) makes it clear that conclusions should be based on the researchers understanding of the problem in context, along with all summary data and tests, and that no single value (being p-values, Bayesian factor or else) can be used support or invalidate a theory.

What to report and how?

Considering that quantitative reports will always have more information content than binary (significant or not) reports, we can always argue that raw and/or normalized effect size, confidence intervals, or Bayes factor must be reported. Reporting everything can however hinder the communication of the main result(s), and we should aim at giving only the information needed, at least in the core of a manuscript. Here I propose to adopt optimal reporting in the result section to keep the message clear, but have detailed supplementary material. When the hypothesis is about the presence/absence or order of an effect, and providing that a study has sufficient power, NHST is appropriate and it is sufficient to report in the text the actual p-value since it conveys the information needed to rule out equivalence. When the hypothesis and/or the discussion involve some quantitative value, and because p-values do not inform on the effect, it is essential to report on effect sizes ( Lakens, 2013 ), preferably accompanied with confidence or credible intervals. The reasoning is simply that one cannot predict and/or discuss quantities without accounting for variability. For the reader to understand and fully appreciate the results, nothing else is needed.

Because science progress is obtained by cumulating evidence ( Rosenthal, 1991 ), scientists should also consider the secondary use of the data. With today’s electronic articles, there are no reasons for not including all of derived data: mean, standard deviations, effect size, CI, Bayes factor should always be included as supplementary tables (or even better also share raw data). It is also essential to report the context in which tests were performed – that is to report all of the tests performed (all t, F, p values) because of the increase type one error rate due to selective reporting (multiple comparisons and p-hacking problems - Ioannidis, 2005 ). Providing all of this information allows (i) other researchers to directly and effectively compare their results in quantitative terms (replication of effects beyond significance, Open Science Collaboration, 2015 ), (ii) to compute power to future studies ( Lakens & Evers, 2014 ), and (iii) to aggregate results for meta-analyses whilst minimizing publication bias ( van Assen et al. , 2014 ).

[version 3; referees: 1 approved

Funding Statement

The author(s) declared that no grants were involved in supporting this work.

  • Christensen R: Testing Fisher, Neyman, Pearson, and Bayes. The American Statistician. 2005; 59 ( 2 ):121–126. 10.1198/000313005X20871 [ CrossRef ] [ Google Scholar ]
  • Cumming G, Maillardet R: Confidence intervals and replication: Where will the next mean fall? Psychological Methods. 2006; 11 ( 3 ):217–227. 10.1037/1082-989X.11.3.217 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dienes Z: Using Bayes to get the most out of non-significant results. Front Psychol. 2014; 5 :781. 10.3389/fpsyg.2014.00781 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fisher RA: Statistical Methods for Research Workers . (Vol. 5th Edition). Edinburgh, UK: Oliver and Boyd.1934. Reference Source [ Google Scholar ]
  • Fisher RA: Statistical Methods and Scientific Induction. Journal of the Royal Statistical Society, Series B. 1955; 17 ( 1 ):69–78. Reference Source [ Google Scholar ]
  • Fisher RA: Statistical methods and scientific inference . (2nd ed.). NewYork: Hafner Publishing,1959. Reference Source [ Google Scholar ]
  • Fisher RA: The Design of Experiments . Hafner Publishing Company, New-York.1971. Reference Source [ Google Scholar ]
  • Frick RW: The appropriate use of null hypothesis testing. Psychol Methods. 1996; 1 ( 4 ):379–390. 10.1037/1082-989X.1.4.379 [ CrossRef ] [ Google Scholar ]
  • Gelman A: P values and statistical practice. Epidemiology. 2013; 24 ( 1 ):69–72. 10.1097/EDE.0b013e31827886f7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Halsey LG, Curran-Everett D, Vowler SL, et al.: The fickle P value generates irreproducible results. Nat Methods. 2015; 12 ( 3 ):179–85. 10.1038/nmeth.3288 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hoekstra R, Morey RD, Rouder JN, et al.: Robust misinterpretation of confidence intervals. Psychon Bull Rev. 2014; 21 ( 5 ):1157–1164. 10.3758/s13423-013-0572-3 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hubbard R, Bayarri MJ: Confusion over measures of evidence (p’s) versus errors ([alpha]’s) in classical statistical testing. The American Statistician. 2003; 57 ( 3 ):171–182. 10.1198/0003130031856 [ CrossRef ] [ Google Scholar ]
  • Ioannidis JP: Why most published research findings are false. PLoS Med. 2005; 2 ( 8 ):e124. 10.1371/journal.pmed.0020124 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Johnson VE: Revised standards for statistical evidence. Proc Natl Acad Sci U S A. 2013; 110 ( 48 ):19313–19317. 10.1073/pnas.1313476110 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Killeen PR: An alternative to null-hypothesis significance tests. Psychol Sci. 2005; 16 ( 5 ):345–353. 10.1111/j.0956-7976.2005.01538.x [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kruschke JK: Bayesian Assessment of Null Values Via Parameter Estimation and Model Comparison. Perspect Psychol Sci. 2011; 6 ( 3 ):299–312. 10.1177/1745691611406925 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Krzywinski M, Altman N: Points of significance: Significance, P values and t -tests. Nat Methods. 2013; 10 ( 11 ):1041–1042. 10.1038/nmeth.2698 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lakens D: Calculating and reporting effect sizes to facilitate cumulative science: a practical primer for t -tests and ANOVAs. Front Psychol. 2013; 4 :863. 10.3389/fpsyg.2013.00863 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lakens D, Evers ER: Sailing From the Seas of Chaos Into the Corridor of Stability: Practical Recommendations to Increase the Informational Value of Studies. Perspect Psychol Sci. 2014; 9 ( 3 ):278–292. 10.1177/1745691614528520 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lindley D: The philosophy of statistics. Journal of the Royal Statistical Society. 2000; 49 ( 3 ):293–337. 10.1111/1467-9884.00238 [ CrossRef ] [ Google Scholar ]
  • Miller J: What is the probability of replicating a statistically significant effect? Psychon Bull Rev. 2009; 16 ( 4 ):617–640. 10.3758/PBR.16.4.617 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Morey RD, Rouder JN: Bayes factor approaches for testing interval null hypotheses. Psychol Methods. 2011; 16 ( 4 ):406–419. 10.1037/a0024377 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Neyman J, Pearson ES: On the Use and Interpretation of Certain Test Criteria for Purposes of Statistical Inference: Part I. Biometrika. 1928; 20A ( 1/2 ):175–240. 10.3389/fpsyg.2015.00245 [ CrossRef ] [ Google Scholar ]
  • Neyman J, Pearson ES: On the problem of the most efficient tests of statistical hypotheses. Philos Trans R Soc Lond Ser A. 1933; 231 ( 694–706 ):289–337. 10.1098/rsta.1933.0009 [ CrossRef ] [ Google Scholar ]
  • Nickerson RS: Null hypothesis significance testing: a review of an old and continuing controversy. Psychol Methods. 2000; 5 ( 2 ):241–301. 10.1037/1082-989X.5.2.241 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nuzzo R: Scientific method: statistical errors. Nature. 2014; 506 ( 7487 ):150–152. 10.1038/506150a [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Open Science Collaboration. PSYCHOLOGY. Estimating the reproducibility of psychological science. Science. 2015; 349 ( 6251 ):aac4716. 10.1126/science.aac4716 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rosenthal R: Cumulating psychology: an appreciation of Donald T. Campbell. Psychol Sci. 1991; 2 ( 4 ):213–221. 10.1111/j.1467-9280.1991.tb00138.x [ CrossRef ] [ Google Scholar ]
  • Savalei V, Dunn E: Is the call to abandon p -values the red herring of the replicability crisis? Front Psychol. 2015; 6 :245. 10.3389/fpsyg.2015.00245 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tan SH, Tan SB: The Correct Interpretation of Confidence Intervals. Proceedings of Singapore Healthcare. 2010; 19 ( 3 ):276–278. 10.1177/201010581001900316 [ CrossRef ] [ Google Scholar ]
  • Turkheimer FE, Aston JA, Cunningham VJ: On the logic of hypothesis testing in functional imaging. Eur J Nucl Med Mol Imaging. 2004; 31 ( 5 ):725–732. 10.1007/s00259-003-1387-7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • van Assen MA, van Aert RC, Nuijten MB, et al.: Why Publishing Everything Is More Effective than Selective Publishing of Statistically Significant Results. PLoS One. 2014; 9 ( 1 ):e84896. 10.1371/journal.pone.0084896 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Walker E, Nowacki AS: Understanding equivalence and noninferiority testing. J Gen Intern Med. 2011; 26 ( 2 ):192–196. 10.1007/s11606-010-1513-8 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wasserstein RL, Lazar NA: The ASA’s Statement on p -Values: Context, Process, and Purpose. The American Statistician. 2016; 70 ( 2 ):129–133. 10.1080/00031305.2016.1154108 [ CrossRef ] [ Google Scholar ]
  • Wilcox R: Introduction to Robust Estimation and Hypothesis Testing . Edition 3, Academic Press, Elsevier: Oxford, UK, ISBN: 978-0-12-386983-8.2012. Reference Source [ Google Scholar ]

Referee response for version 3

Dorothy vera margaret bishop.

1 Department of Experimental Psychology, University of Oxford, Oxford, UK

I can see from the history of this paper that the author has already been very responsive to reviewer comments, and that the process of revising has now been quite protracted.

That makes me reluctant to suggest much more, but I do see potential here for making the paper more impactful. So my overall view is that, once a few typos are fixed (see below), this could be published as is, but I think there is an issue with the potential readership and that further revision could overcome this.

I suspect my take on this is rather different from other reviewers, as I do not regard myself as a statistics expert, though I am on the more quantitative end of the continuum of psychologists and I try to keep up to date. I think I am quite close to the target readership , insofar as I am someone who was taught about statistics ages ago and uses stats a lot, but never got adequate training in the kinds of topic covered by this paper. The fact that I am aware of controversies around the interpretation of confidence intervals etc is simply because I follow some discussions of this on social media. I am therefore very interested to have a clear account of these issues.

This paper contains helpful information for someone in this position, but it is not always clear, and I felt the relevance of some of the content was uncertain. So here are some recommendations:

  • As one previous reviewer noted, it’s questionable that there is a need for a tutorial introduction, and the limited length of this article does not lend itself to a full explanation. So it might be better to just focus on explaining as clearly as possible the problems people have had in interpreting key concepts. I think a title that made it clear this was the content would be more appealing than the current one.
  • P 3, col 1, para 3, last sentence. Although statisticians always emphasise the arbitrary nature of p < .05, we all know that in practice authors who use other values are likely to have their analyses queried. I wondered whether it would be useful here to note that in some disciplines different cutoffs are traditional, e.g. particle physics. Or you could cite David Colquhoun’s paper in which he recommends using p < .001 ( http://rsos.royalsocietypublishing.org/content/1/3/140216) - just to be clear that the traditional p < .05 has been challenged.

What I can’t work out is how you would explain the alpha from Neyman-Pearson in the same way (though I can see from Figure 1 that with N-P you could test an alternative hypothesis, such as the idea that the coin would be heads 75% of the time).

‘By failing to reject, we simply continue to assume that H0 is true, which implies that one cannot….’ have ‘In failing to reject, we do not assume that H0 is true; one cannot argue against a theory from a non-significant result.’

I felt most readers would be interested to read about tests of equivalence and Bayesian approaches, but many would be unfamiliar with these and might like to see an example of how they work in practice – if space permitted.

  • Confidence intervals: I simply could not understand the first sentence – I wondered what was meant by ‘builds’ here. I understand about difficulties in comparing CI across studies when sample sizes differ, but I did not find the last sentence on p 4 easy to understand.
  • P 5: The sentence starting: ‘The alpha value has the same interpretation’ was also hard to understand, especially the term ‘1-alpha CI’. Here too I felt some concrete illustration might be helpful to the reader. And again, I also found the reference to Bayesian intervals tantalising – I think many readers won’t know how to compute these and something like a figure comparing a traditional CI with a Bayesian interval and giving a source for those who want to read on would be very helpful. The reference to ‘credible intervals’ in the penultimate paragraph is very unclear and needs a supporting reference – most readers will not be familiar with this concept.

P 3, col 1, para 2, line 2; “allows us to compute”

P 3, col 2, para 2, ‘probability of replicating’

P 3, col 2, para 2, line 4 ‘informative about’

P 3, col 2, para 4, line 2 delete ‘of’

P 3, col 2, para 5, line 9 – ‘conditioned’ is either wrong or too technical here: would ‘based’ be acceptable as alternative wording

P 3, col 2, para 5, line 13 ‘This dichotomisation allows one to distinguish’

P 3, col 2, para 5, last sentence, delete ‘Alternatively’.

P 3, col 2, last para line 2 ‘first’

P 4, col 2, para 2, last sentence is hard to understand; not sure if this is better: ‘If sample sizes differ between studies, the distribution of CIs cannot be specified a priori’

P 5, col 1, para 2, ‘a pattern of order’ – I did not understand what was meant by this

P 5, col 1, para 2, last sentence unclear: possible rewording: “If the goal is to test the size of an effect then NHST is not the method of choice, since testing can only reject the null hypothesis.’ (??)

P 5, col 1, para 3, line 1 delete ‘that’

P 5, col 1, para 3, line 3 ‘on’ -> ‘by’

P 5, col 2, para 1, line 4 , rather than ‘Here I propose to adopt’ I suggest ‘I recommend adopting’

P 5, col 2, para 1, line 13 ‘with’ -> ‘by’

P 5, col 2, para 1 – recommend deleting last sentence

P 5, col 2, para 2, line 2 ‘consider’ -> ‘anticipate’

P 5, col 2, para 2, delete ‘should always be included’

P 5, col 2, para 2, ‘type one’ -> ‘Type I’

I have read this submission. I believe that I have an appropriate level of expertise to confirm that it is of an acceptable scientific standard, however I have significant reservations, as outlined above.

The University of Edinburgh, UK

I wondered about changing the focus slightly and modifying the title to reflect this to say something like: Null hypothesis significance testing: a guide to commonly misunderstood concepts and recommendations for good practice

Thank you for the suggestion – you indeed saw the intention behind the ‘tutorial’ style of the paper.

  • P 3, col 1, para 3, last sentence. Although statisticians always emphasise the arbitrary nature of p < .05, we all know that in practice authors who use other values are likely to have their analyses queried. I wondered whether it would be useful here to note that in some disciplines different cutoffs are traditional, e.g. particle physics. Or you could cite David Colquhoun’s paper in which he recommends using p < .001 ( http://rsos.royalsocietypublishing.org/content/1/3/140216)  - just to be clear that the traditional p < .05 has been challenged.

I have added a sentence on this citing Colquhoun 2014 and the new Benjamin 2017 on using .005.

I agree that this point is always hard to appreciate, especially because it seems like in practice it makes little difference. I added a paragraph but using reaction times rather than a coin toss – thanks for the suggestion.

Added an example based on new table 1, following figure 1 – giving CI, equivalence tests and Bayes Factor (with refs to easy to use tools)

Changed builds to constructs (this simply means they are something we build) and added that the implication that probability coverage is not warranty when sample size change, is that we cannot compare CI.

I changed ‘ i.e. we accept that 1-alpha CI are wrong in alpha percent of the times in the long run’ to ‘, ‘e.g. a 95% CI is wrong in 5% of the times in the long run (i.e. if we repeat the experiment many times).’ – for Bayesian intervals I simply re-cited Morey & Rouder, 2011.

It is not the CI cannot be specified, it’s that the interval is not predictive of anything anymore! I changed it to ‘If sample sizes, however, differ between studies, there is no warranty that a CI from one study will be true at the rate alpha in a different study, which implies that CI cannot be compared across studies at this is rarely the same sample sizes’

I added (i.e. establish that A > B) – we test that conditions are ordered, but without further specification of the probability of that effect nor its size

Yes it works – thx

P 5, col 2, para 2, ‘type one’ -> ‘Type I’ 

Typos fixed, and suggestions accepted – thanks for that.

Stephen J. Senn

1 Luxembourg Institute of Health, Strassen, L-1445, Luxembourg

The revisions are OK for me, and I have changed my status to Approved.

I have read this submission. I believe that I have an appropriate level of expertise to confirm that it is of an acceptable scientific standard.

Referee response for version 2

On the whole I think that this article is reasonable, my main reservation being that I have my doubts on whether the literature needs yet another tutorial on this subject.

A further reservation I have is that the author, following others, stresses what in my mind is a relatively unimportant distinction between the Fisherian and Neyman-Pearson (NP) approaches. The distinction stressed by many is that the NP approach leads to a dichotomy accept/reject based on probabilities established in advance, whereas the Fisherian approach uses tail area probabilities calculated from the observed statistic. I see this as being unimportant and not even true. Unless one considers that the person carrying out a hypothesis test (original tester) is mandated to come to a conclusion on behalf of all scientific posterity, then one must accept that any remote scientist can come to his or her conclusion depending on the personal type I error favoured. To operate the results of an NP test carried out by the original tester, the remote scientist then needs to know the p-value. The type I error rate is then compared to this to come to a personal accept or reject decision (1). In fact Lehmann (2), who was an important developer of and proponent of the NP system, describes exactly this approach as being good practice. (See Testing Statistical Hypotheses, 2nd edition P70). Thus using tail-area probabilities calculated from the observed statistics does not constitute an operational difference between the two systems.

A more important distinction between the Fisherian and NP systems is that the former does not use alternative hypotheses(3). Fisher's opinion was that the null hypothesis was more primitive than the test statistic but that the test statistic was more primitive than the alternative hypothesis. Thus, alternative hypotheses could not be used to justify choice of test statistic. Only experience could do that.

Further distinctions between the NP and Fisherian approach are to do with conditioning and whether a null hypothesis can ever be accepted.

I have one minor quibble about terminology. As far as I can see, the author uses the usual term 'null hypothesis' and the eccentric term 'nil hypothesis' interchangeably. It would be simpler if the latter were abandoned.

Referee response for version 1

Marcel alm van assen.

1 Department of Methodology and Statistics, Tilburgh University, Tilburg, Netherlands

Null hypothesis significance testing (NHST) is a difficult topic, with misunderstandings arising easily. Many texts, including basic statistics books, deal with the topic, and attempt to explain it to students and anyone else interested. I would refer to a good basic text book, for a detailed explanation of NHST, or to a specialized article when wishing an explaining the background of NHST. So, what is the added value of a new text on NHST? In any case, the added value should be described at the start of this text. Moreover, the topic is so delicate and difficult that errors, misinterpretations, and disagreements are easy. I attempted to show this by giving comments to many sentences in the text.

Abstract: “null hypothesis significance testing is the statistical method of choice in biological, biomedical and social sciences to investigate if an effect is likely”. No, NHST is the method to test the hypothesis of no effect.

Intro: “Null hypothesis significance testing (NHST) is a method of statistical inference by which an observation is tested against a hypothesis of no effect or no relationship.” What is an ‘observation’? NHST is difficult to describe in one sentence, particularly here. I would skip this sentence entirely, here.

Section on Fisher; also explain the one-tailed test.

Section on Fisher; p(Obs|H0) does not reflect the verbal definition (the ‘or more extreme’ part).

Section on Fisher; use a reference and citation to Fisher’s interpretation of the p-value

Section on Fisher; “This was however only intended to be used as an indication that there is something in the data that deserves further investigation. The reason for this is that only H0 is tested whilst the effect under study is not itself being investigated.” First sentence, can you give a reference? Many people say a lot about Fisher’s intentions, but the good man is dead and cannot reply… Second sentence is a bit awkward, because the effect is investigated in a way, by testing the H0.

Section on p-value; Layout and structure can be improved greatly, by first again stating what the p-value is, and then statement by statement, what it is not, using separate lines for each statement. Consider adding that the p-value is randomly distributed under H0 (if all the assumptions of the test are met), and that under H1 the p-value is a function of population effect size and N; the larger each is, the smaller the p-value generally is.

Skip the sentence “If there is no effect, we should replicate the absence of effect with a probability equal to 1-p”. Not insightful, and you did not discuss the concept ‘replicate’ (and do not need to).

Skip the sentence “The total probability of false positives can also be obtained by aggregating results ( Ioannidis, 2005 ).” Not strongly related to p-values, and introduces unnecessary concepts ‘false positives’ (perhaps later useful) and ‘aggregation’.

Consider deleting; “If there is an effect however, the probability to replicate is a function of the (unknown) population effect size with no good way to know this from a single experiment ( Killeen, 2005 ).”

The following sentence; “ Finally, a (small) p-value  is not an indication favouring a hypothesis . A low p-value indicates a misfit of the null hypothesis to the data and cannot be taken as evidence in favour of a specific alternative hypothesis more than any other possible alternatives such as measurement error and selection bias ( Gelman, 2013 ).” is surely not mainstream thinking about NHST; I would surely delete that sentence. In NHST, a p-value is used for testing the H0. Why did you not yet discuss significance level? Yes, before discussing what is not a p-value, I would explain NHST (i.e., what it is and how it is used). 

Also the next sentence “The more (a priori) implausible the alternative hypothesis, the greater the chance that a finding is a false alarm ( Krzywinski & Altman, 2013 ;  Nuzzo, 2014 ).“ is not fully clear to me. This is a Bayesian statement. In NHST, no likelihoods are attributed to hypotheses; the reasoning is “IF H0 is true, then…”.

Last sentence: “As  Nickerson (2000)  puts it ‘theory corroboration requires the testing of multiple predictions because the chance of getting statistically significant results for the wrong reasons in any given case is high’.” What is relation of this sentence to the contents of this section, precisely?

Next section: “For instance, we can estimate that the probability of a given F value to be in the critical interval [+2 +∞] is less than 5%” This depends on the degrees of freedom.

“When there is no effect (H0 is true), the erroneous rejection of H0 is known as type I error and is equal to the p-value.” Strange sentence. The Type I error is the probability of erroneously rejecting the H0 (so, when it is true). The p-value is … well, you explained it before; it surely does not equal the Type I error.

Consider adding a figure explaining the distinction between Fisher’s logic and that of Neyman and Pearson.

“When the test statistics falls outside the critical region(s)” What is outside?

“There is a profound difference between accepting the null hypothesis and simply failing to reject it ( Killeen, 2005 )” I agree with you, but perhaps you may add that some statisticians simply define “accept H0’” as obtaining a p-value larger than the significance level. Did you already discuss the significance level, and it’s mostly used values?

“To accept or reject equally the null hypothesis, Bayesian approaches ( Dienes, 2014 ;  Kruschke, 2011 ) or confidence intervals must be used.” Is ‘reject equally’ appropriate English? Also using Cis, one cannot accept the H0.

Do you start discussing alpha only in the context of Cis?

“CI also indicates the precision of the estimate of effect size, but unless using a percentile bootstrap approach, they require assumptions about distributions which can lead to serious biases in particular regarding the symmetry and width of the intervals ( Wilcox, 2012 ).” Too difficult, using new concepts. Consider deleting.

“Assuming the CI (a)symmetry and width are correct, this gives some indication about the likelihood that a similar value can be observed in future studies, with 95% CI giving about 83% chance of replication success ( Lakens & Evers, 2014 ).” This statement is, in general, completely false. It very much depends on the sample sizes of both studies. If the replication study has a much, much, much larger N, then the probability that the original CI will contain the effect size of the replication approaches (1-alpha)*100%. If the original study has a much, much, much larger N, then the probability that the original Ci will contain the effect size of the replication study approaches 0%.

“Finally, contrary to p-values, CI can be used to accept H0. Typically, if a CI includes 0, we cannot reject H0. If a critical null region is specified rather than a single point estimate, for instance [-2 +2] and the CI is included within the critical null region, then H0 can be accepted. Importantly, the critical region must be specified a priori and cannot be determined from the data themselves.” No. H0 cannot be accepted with Cis.

“The (posterior) probability of an effect can however not be obtained using a frequentist framework.” Frequentist framework? You did not discuss that, yet.

“X% of times the CI obtained will contain the same parameter value”. The same? True, you mean?

“e.g. X% of the times the CI contains the same mean” I do not understand; which mean?

“The alpha value has the same interpretation as when using H0, i.e. we accept that 1-alpha CI are wrong in alpha percent of the times. “ What do you mean, CI are wrong? Consider rephrasing.

“To make a statement about the probability of a parameter of interest, likelihood intervals (maximum likelihood) and credibility intervals (Bayes) are better suited.” ML gives the likelihood of the data given the parameter, not the other way around.

“Many of the disagreements are not on the method itself but on its use.” Bayesians may disagree.

“If the goal is to establish the likelihood of an effect and/or establish a pattern of order, because both requires ruling out equivalence, then NHST is a good tool ( Frick, 1996 )” NHST does not provide evidence on the likelihood of an effect.

“If the goal is to establish some quantitative values, then NHST is not the method of choice.” P-values are also quantitative… this is not a precise sentence. And NHST may be used in combination with effect size estimation (this is even recommended by, e.g., the American Psychological Association (APA)).

“Because results are conditioned on H0, NHST cannot be used to establish beliefs.” It can reinforce some beliefs, e.g., if H0 or any other hypothesis, is true.

“To estimate the probability of a hypothesis, a Bayesian analysis is a better alternative.” It is the only alternative?

“Note however that even when a specific quantitative prediction from a hypothesis is shown to be true (typically testing H1 using Bayes), it does not prove the hypothesis itself, it only adds to its plausibility.” How can we show something is true?

I do not agree on the contents of the last section on ‘minimal reporting’. I prefer ‘optimal reporting’ instead, i.e., the reporting the information that is essential to the interpretation of the result, to any ready, which may have other goals than the writer of the article. This reporting includes, for sure, an estimate of effect size, and preferably a confidence interval, which is in line with recommendations of the APA.

I have read this submission. I believe that I have an appropriate level of expertise to state that I do not consider it to be of an acceptable scientific standard, for reasons outlined above.

The idea of this short review was to point to common interpretation errors (stressing again and again that we are under H0) being in using p-values or CI, and also proposing reporting practices to avoid bias. This is now stated at the end of abstract.

Regarding text books, it is clear that many fail to clearly distinguish Fisher/Pearson/NHST, see Glinet et al (2012) J. Exp Education 71, 83-92. If you have 1 or 2 in mind that you know to be good, I’m happy to include them.

I agree – yet people use it to investigate (not test) if an effect is likely. The issue here is wording. What about adding this distinction at the end of the sentence?: ‘null hypothesis significance testing is the statistical method of choice in biological, biomedical and social sciences used to investigate if an effect is likely, even though it actually tests for the hypothesis of no effect’.

I think a definition is needed, as it offers a starting point. What about the following: ‘NHST is a method of statistical inference by which an experimental factor is tested against a hypothesis of no effect or no relationship based on a given observation’

The section on Fisher has been modified (more or less) as suggested: (1) avoiding talking about one or two tailed tests (2) updating for p(Obs≥t|H0) and (3) referring to Fisher more explicitly (ie pages from articles and book) ; I cannot tell his intentions but these quotes leave little space to alternative interpretations.

The reasoning here is as you state yourself, part 1: ‘a p-value is used for testing the H0; and part 2: ‘no likelihoods are attributed to hypotheses’ it follows we cannot favour a hypothesis. It might seems contentious but this is the case that all we can is to reject the null – how could we favour a specific alternative hypothesis from there? This is explored further down the manuscript (and I now point to that) – note that we do not need to be Bayesian to favour a specific H1, all I’m saying is this cannot be attained with a p-value.

The point was to emphasise that a p value is not there to tell us a given H1 is true and can only be achieved through multiple predictions and experiments. I deleted it for clarity.

This sentence has been removed

Indeed, you are right and I have modified the text accordingly. When there is no effect (H0 is true), the erroneous rejection of H0 is known as type 1 error. Importantly, the type 1 error rate, or alpha value is determined a priori. It is a common mistake but the level of significance (for a given sample) is not the same as the frequency of acceptance alpha found on repeated sampling (Fisher, 1955).

A figure is now presented – with levels of acceptance, critical region, level of significance and p-value.

I should have clarified further here – as I was having in mind tests of equivalence. To clarify, I simply states now: ‘To accept the null hypothesis, tests of equivalence or Bayesian approaches must be used.’

It is now presented in the paragraph before.

Yes, you are right, I completely overlooked this problem. The corrected sentence (with more accurate ref) is now “Assuming the CI (a)symmetry and width are correct, this gives some indication about the likelihood that a similar value can be observed in future studies. For future studies of the same sample size, 95% CI giving about 83% chance of replication success (Cumming and Mallardet, 2006). If sample sizes differ between studies, CI do not however warranty any a priori coverage”.

Again, I had in mind equivalence testing, but in both cases you are right we can only reject and I therefore removed that sentence.

Yes, p-values must be interpreted in context with effect size, but this is not what people do. The point here is to be pragmatic, does and don’t. The sentence was changed.

Not for testing, but for probability, I am not aware of anything else.

Cumulative evidence is, in my opinion, the only way to show it. Even in hard science like physics multiple experiments. In the recent CERN study on finding Higgs bosons, 2 different and complementary experiments ran in parallel – and the cumulative evidence was taken as a proof of the true existence of Higgs bosons.

Daniel Lakens

1 School of Innovation Sciences, Eindhoven University of Technology, Eindhoven, Netherlands

I appreciate the author's attempt to write a short tutorial on NHST. Many people don't know how to use it, so attempts to educate people are always worthwhile. However, I don't think the current article reaches it's aim. For one, I think it might be practically impossible to explain a lot in such an ultra short paper - every section would require more than 2 pages to explain, and there are many sections. Furthermore, there are some excellent overviews, which, although more extensive, are also much clearer (e.g., Nickerson, 2000 ). Finally, I found many statements to be unclear, and perhaps even incorrect (noted below). Because there is nothing worse than creating more confusion on such a topic, I have extremely high standards before I think such a short primer should be indexed. I note some examples of unclear or incorrect statements below. I'm sorry I can't make a more positive recommendation.

“investigate if an effect is likely” – ambiguous statement. I think you mean, whether the observed DATA is probable, assuming there is no effect?

The Fisher (1959) reference is not correct – Fischer developed his method much earlier.

“This p-value thus reflects the conditional probability of achieving the observed outcome or larger, p(Obs|H0)” – please add 'assuming the null-hypothesis is true'.

“p(Obs|H0)” – explain this notation for novices.

“Following Fisher, the smaller the p-value, the greater the likelihood that the null hypothesis is false.”  This is wrong, and any statement about this needs to be much more precise. I would suggest direct quotes.

“there is something in the data that deserves further investigation” –unclear sentence.

“The reason for this” – unclear what ‘this’ refers to.

“ not the probability of the null hypothesis of being true, p(H0)” – second of can be removed?

“Any interpretation of the p-value in relation to the effect under study (strength, reliability, probability) is indeed

wrong, since the p-value is conditioned on H0”  - incorrect. A big problem is that it depends on the sample size, and that the probability of a theory depends on the prior.

“If there is no effect, we should replicate the absence of effect with a probability equal to 1-p.” I don’t understand this, but I think it is incorrect.

“The total probability of false positives can also be obtained by aggregating results (Ioannidis, 2005).” Unclear, and probably incorrect.

“By failing to reject, we simply continue to assume that H0 is true, which implies that one cannot, from a nonsignificant result, argue against a theory” – according to which theory? From a NP perspective, you can ACT as if the theory is false.

“(Lakens & Evers, 2014”) – we are not the original source, which should be cited instead.

“ Typically, if a CI includes 0, we cannot reject H0.”  - when would this not be the case? This assumes a CI of 1-alpha.

“If a critical null region is specified rather than a single point estimate, for instance [-2 +2] and the CI is included within the critical null region, then H0 can be accepted.” – you mean practically, or formally? I’m pretty sure only the former.

The section on ‘The (correct) use of NHST’ seems to conclude only Bayesian statistics should be used. I don’t really agree.

“ we can always argue that effect size, power, etc. must be reported.” – which power? Post-hoc power? Surely not? Other types are unknown. So what do you mean?

The recommendation on what to report remains vague, and it is unclear why what should be reported.

This sentence was changed, following as well the other reviewer, to ‘null hypothesis significance testing is the statistical method of choice in biological, biomedical and social sciences to investigate if an effect is likely, even though it actually tests whether the observed data are probable, assuming there is no effect’

Changed, refers to Fisher 1925

I changed a little the sentence structure, which should make explicit that this is the condition probability.

This has been changed to ‘[…] to decide whether the evidence is worth additional investigation and/or replication (Fisher, 1971 p13)’

my mistake – the sentence structure is now ‘ not the probability of the null hypothesis p(H0), of being true,’ ; hope this makes more sense (and this way refers back to p(Obs>t|H0)

Fair enough – my point was to stress the fact that p value and effect size or H1 have very little in common, but yes that the part in common has to do with sample size. I left the conditioning on H0 but also point out the dependency on sample size.

The whole paragraph was changed to reflect a more philosophical take on scientific induction/reasoning. I hope this is clearer.

Changed to refer to equivalence testing

I rewrote this, as to show frequentist analysis can be used  - I’m trying to sell Bayes more than any other approach.

I’m arguing we should report it all, that’s why there is no exhausting list – I can if needed.

null hypothesis confidence interval

Random Error

  •   1  
  • |   2  
  • |   3  
  • |   4  
  • |   5  
  • |   6  
  • |   7  

On This Page sidebar

Confidence Intervals and p-Values

The importance of precision, with "non-significant" results, with "significant" results.

Learn More sidebar

Epi_Tools.XLSX

All Modules

Confidence intervals are calculated from the same equations that generate p-values, so, not surprisingly, there is a relationship between the two, and confidence intervals for measures of association are often used to address the question of "statistical significance" even if a p-value is not calculated. We already noted that one way of stating the null hypothesis is to state that a risk ratio or an odds ratio is 1.0. We also noted that the point estimate is the most likely value, based on the observed data, and the 95% confidence interval quantifies the random error associated with our estimate, and it can also be interpreted as the range within which the true value is likely to lie with 95% confidence. This means that values outside the 95% confidence interval are unlikely to be the true value. Therefore, if the null value (RR=1.0 or OR=1.0) is not contained within the 95% confidence interval, then the probability that the null is the true value is less than 5%. Conversely, if the null is contained within the 95% confidence interval, then the null is one of the values that is consistent with the observed data, so the null hypothesis cannot be rejected.  

NOTE: Such a usage is unfortunate in my view because it is essentially using a confidence interval to make an accept/reject decision rather than focusing on it as a measure of precision, and it focuses all attention on one side of a two-sided measure (for example, if the upper and lower limits of a confidence interval are .90 and 2.50, there is just as great a chance that the true result is 2.50 as .90).

An easy way to remember the relationship between a 95% confidence interval and a p-value of 0.05 is to think of the confidence interval as arms that "embrace" values that are consistent with the data. If the null value is "embraced", then it is certainly not rejected , i.e. the p-value must be greater than 0.05 (not statistically significant) if the null value is within the interval. However, if the 95% CI excludes the null value, then the null hypothesis has been rejected , and the p-value must be < 0.05.

If the null value, meaning no difference is within the 95% confidence interval (i.e. embraced by it) then the results are not statistically significant.

Video Summary: Confidence Intervals for Risk Ratio, Odds Ratio, and Rate Ratio (8:35)

Link to a transcrip of the video

The difference between the perspective provided by the confidence interval and significance testing is particularly clear when considering non-significant results. The image below shows two confidence intervals; neither of them is "statistically significant" using the criterion of P < 0.05, because both of them embrace the null (risk ratio = 1.0). However, one should view these two estimates differently. The estimate with the wide confidence interval was likely obtained with a small sample size and a lot of potential for random error. However, even though it is not statistically significant, the point estimate (i.e., the estimated risk ratio or odds ratio) was somewhere around four, raising the possibility of an important effect. In this case one might want to explore this further by repeating the study with a larger sample size. Repeating the study with a larger sample would certainly not guarantee a statistically significant result, but it would provide a more precise estimate. The other estimate that is depicted is also non-significant, but it is a much narrower, i.e., more precise estimate, and we are confident that the true value is likely to be close to the null value. Even if there were a difference between the groups, it is likely to be a very small difference that may have little if any clinical significance. So, in this case, one would not be inclined to repeat the study.

A narrow confidence interval and a wide confidence interval both include the null value. The greater precision of the larger sample that produced the narrow confidence interval indicates that it is unlikely that there is a clinically important effect, but the study with the wide confidence interval should perhaps be repeated with a larger sample in order to avoid missing a clinically important effect.

For example, even if a huge study were undertaken that indicated a risk ratio of 1.03 with a 95% confidence interval of 1.02 - 1.04, this would indicate an increase in risk of only 2 - 4%. Even if this were true, it would not be important, and it might very well still be the result of biases or residual confounding. Consequently, the narrow confidence interval provides strong evidence that there is little or no association.

The next figure illustrates two study results that are both statistically significant at P < 0.05, because both confidence intervals lie entirely above the null value (RR or OR = 1). The upper result has a point estimate of about two, and its confidence interval ranges from about 0.5 to 3.0, and the lower result shows a point estimate of about 6 with a confidence interval that ranges from 0.5 to about 12. The narrower, more precise estimate enables us to be confident that there is about a two-fold increase in risk among those who have the exposure of interest. In contrast, the study with the wide confidence interval is "statistically significant," but it leaves us uncertain about the magnitude of the effect. Is the increase in risk relatively modest or is it huge? We just don't know.

A narrow and a wide confidence interval. Both lie completely above the null value, which is risk ratio=1

So, regardless of whether a study's results meet the criterion for statistically significance, a more important consideration is the precision of the estimate.

   

Statistical Significance Using Confidence Intervals. Drag the card from the bottom to the correct category.

return to top | previous page | next page

Click to close

Content ©2016. All Rights Reserved. Date last modified: June 16, 2016. Wayne W. LaMorte, MD, PhD, MPH,

Using a confidence interval to decide whether to reject the null hypothesis

Suppose that you do a hypothesis test. Remember that the decision to reject the null hypothesis (H 0 ) or fail to reject it can be based on the p-value and your chosen significance level (also called α). If the p-value is less than or equal to α, you reject H 0 ; if it is greater than α, you fail to reject H 0 .

  • If the reference value specified in H 0 lies outside the interval (that is, is less than the lower bound or greater than the upper bound), you can reject H 0 .
  • If the reference value specified in H 0 lies within the interval (that is, is not less than the lower bound or greater than the upper bound), you fail to reject H 0 .
  • Minitab.com
  • License Portal
  • Cookie Settings

You are now leaving support.minitab.com.

Click Continue to proceed to:

Rejecting the Null Hypothesis Using Confidence Intervals

  • Tech Trends

Rejecting the Null Hypothesis Using Confidence Intervals

In an introductory statistics class, there are three main topics that are taught: descriptive statistics and data visualizations, probability and sampling distributions, and statistical inference. Within statistical inference, there are two key methods of statistical inference that are taught, viz. confidence intervals and hypothesis testing . While these two methods are always taught when learning data science and related fields, it is rare that the relationship between these two methods is properly elucidated.

In this article, we’ll begin by defining and describing each method of statistical inference in turn and along the way, state what statistical inference is, and perhaps more importantly, what it isn’t. Then we’ll describe the relationship between the two. While it is typically the case that confidence intervals are taught before hypothesis testing when learning statistics, we’ll begin with the latter since it will allow us to define statistical significance.

Hypothesis Tests

The purpose of a hypothesis test is to answer whether random chance might be responsible for an observed effect. Hypothesis tests use sample statistics to test a hypothesis about population parameters. The null hypothesis, H 0 , is a statement that represents the assumed status quo regarding a variable or variables and it is always about a population characteristic. Some of the ways the null hypothesis is typically glossed are: the population variable is equal to a particular value or there is no difference between the population variables . For example:

  • H 0 : μ = 61 in (The mean height of the population of American men is 69 inches)
  • H 0 : p 1 -p 2 = 0 (The difference in the population proportions of women who prefer football over baseball and the population proportion of men who prefer football over baseball is 0.)

Note that the null hypothesis always has the equal sign.

The alternative hypothesis, denoted either H 1 or H a , is the statement that is opposed to the null hypothesis (e.g., the population variable is not equal to a particular value  or there is a difference between the population variables ):

  • H 1 : μ > 61 im (The mean height of the population of American men is greater than 69 inches.)
  • H 1 : p 1 -p 2 ≠ 0 (The difference in the population proportions of women who prefer football over baseball and the population proportion of men who prefer football over baseball is not 0.)

The alternative hypothesis is typically the claim that the researcher hopes to show and it always contains the strict inequality symbols (‘<’ left-sided or left-tailed, ‘≠’ two-sided or two-tailed, and ‘>’ right-sided or right-tailed).

When carrying out a test of H 0 vs. H 1 , the null hypothesis H 0 will be rejected in favor of the alternative hypothesis only if the sample provides convincing evidence that H 0 is false. As such, a statistical hypothesis test is only capable of demonstrating strong support for the alternative hypothesis by rejecting the null hypothesis.

When the null hypothesis is not rejected, it does not mean that there is strong support for the null hypothesis (since it was assumed to be true); rather, only that there is not convincing evidence against the null hypothesis. As such, we never use the phrase “accept the null hypothesis.”

In the classical method of performing hypothesis testing, one would have to find what is called the test statistic and use a table to find the corresponding probability. Happily, due to the advancement of technology, one can use Python (as is done in the Flatiron’s Data Science Bootcamp ) and get the required value directly using a Python library like stats models . This is the p-value , which is short for the probability value.

The p-value is a measure of inconsistency between the hypothesized value for a population characteristic and the observed sample. The p -value is the probability, under the assumption the null hypothesis is true, of obtaining a test statistic value that is a measure of inconsistency between the null hypothesis and the data. If the p -value is less than or equal to the probability of the Type I error, then we can reject the null hypothesis and we have sufficient evidence to support the alternative hypothesis.

Typically the probability of a Type I error ɑ, more commonly known as the level of significance , is set to be 0.05, but it is often prudent to have it set to values less than that such as 0.01 or 0.001. Thus, if p -value ≤ ɑ, then we reject the null hypothesis and we interpret this as saying there is a statistically significant difference between the sample and the population. So if the p -value=0.03 ≤ 0.05 = ɑ, then we would reject the null hypothesis and so have statistical significance, whereas if p -value=0.08 ≥ 0.05 = ɑ, then we would fail to reject the null hypothesis and there would not be statistical significance.

Confidence Intervals

The other primary form of statistical inference are confidence intervals. While hypothesis tests are concerned with testing a claim, the purpose of a confidence interval is to estimate an unknown population characteristic. A confidence interval is an interval of plausible values for a population characteristic. They are constructed so that we have a chosen level of confidence that the actual value of the population characteristic will be between the upper and lower endpoints of the open interval.

The structure of an individual confidence interval is the sample estimate of the variable of interest margin of error. The margin of error is the product of a multiplier value and the standard error, s.e., which is based on the standard deviation and the sample size. The multiplier is where the probability, of level of confidence, is introduced into the formula.

The confidence level is the success rate of the method used to construct a confidence interval. A confidence interval estimating the proportion of American men who state they are an avid fan of the NFL could be (0.40, 0.60) with a 95% level of confidence. The level of confidence is not the probability that that population characteristic is in the confidence interval, but rather refers to the method that is used to construct the confidence interval.

For example, a 95% confidence interval would be interpreted as if one constructed 100 confidence intervals, then 95 of them would contain the true population characteristic. 

Errors and Power

A Type I error, or a false positive, is the error of finding a difference that is not there, so it is the probability of incorrectly rejecting a true null hypothesis is ɑ, where ɑ is the level of significance. It follows that the probability of correctly failing to reject a true null hypothesis is the complement of it, viz. 1 – ɑ. For a particular hypothesis test, if ɑ = 0.05, then its complement would be 0.95 or 95%.

While we are not going to expand on these ideas, we note the following two related probabilities. A Type II error, or false negative, is the probability of failing to reject a false null hypothesis where the probability of a type II error is β and the power is the probability of correctly rejecting a false null hypothesis where power = 1 – β. In common statistical practice, one typically only speaks of the level of significance and the power.

The following table summarizes these ideas , where the column headers refer to what is actually the case, but is unknown. (If the truth or falsity of the null value was truly known, we wouldn’t have to do statistics.)

null hypothesis confidence interval

Hypothesis Tests and Confidence Intervals

Since hypothesis tests and confidence intervals are both methods of statistical inference, then it is reasonable to wonder if they are equivalent in some way. The answer is yes, which means that we can perform hypothesis testing using confidence intervals.

Returning to the example where we have an estimate of the proportion of American men that are avid fans of the NFL, we had (0.40, 0.60) at a 95% confidence level. As a hypothesis test, we could have the alternative hypothesis as H 1 ≠ 0.51. Since the null value of 0.51 lies within the confidence interval, then we would fail to reject the null hypothesis at ɑ = 0.05.

On the other hand, if H 1 ≠ 0.61, then since 0.61 is not in the confidence interval we can reject the null hypothesis at ɑ = 0.05. Note that the confidence level of 95% and the level of significance at ɑ = 0.05 = 5%  are complements, which is the “H o is True” column in the above table.

In general, one can reject the null hypothesis given a null value and a confidence interval for a two-sided test if the null value is not in the confidence interval where the confidence level and level of significance are complements. For one-sided tests, one can still perform a hypothesis test with the confidence level and null value. Not only is there an added layer of complexity for this equivalence, it is the best practice to perform two-sided hypothesis tests since one is not prejudicing the direction of the alternative.

In this discussion of hypothesis testing and confidence intervals, we not only understand when these two methods of statistical inference can be equivalent, but now have a deeper understanding of statistical significance itself and therefore, statistical inference.

Learn More About Data Science at Flatiron

The curriculum in our Data Science Bootcamp incorporates the latest technologies, including artificial intelligence (AI) tools. Download the syllabus to see what you can learn, or book a 10-minute call with Admissions to learn about full-time and part-time attendance opportunities.

null hypothesis confidence interval

About Brendan Patrick Purdy

Brendan is the senior curriculum developer for data science at the Flatiron School. He holds degrees in mathematics, data science, and philosophy, and enjoys modeling neural networks with the Python library TensorFlow.

Related Resources

TR NYC Tour

NYC Campus Tour

null hypothesis confidence interval

Quantifying Rafael Nadal’s Dominance with French Open Data

The Art of Data Exploration

The Art of Data Exploration

Privacy overview.

CookieDurationDescription
cookielawinfo-checkbox-analytics11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Analytics".
cookielawinfo-checkbox-functional11 monthsThe cookie is set by GDPR cookie consent to record the user consent for the cookies in the category "Functional".
cookielawinfo-checkbox-necessary11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookies is used to store the user consent for the cookies in the category "Necessary".
cookielawinfo-checkbox-others11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Other.
cookielawinfo-checkbox-performance11 monthsThis cookie is set by GDPR Cookie Consent plugin. The cookie is used to store the user consent for the cookies in the category "Performance".
viewed_cookie_policy11 monthsThe cookie is set by the GDPR Cookie Consent plugin and is used to store whether or not user has consented to the use of cookies. It does not store any personal data.

Save 10% on All AnalystPrep 2024 Study Packages with Coupon Code BLOG10 .

  • Payment Plans
  • Product List
  • Partnerships

AnalystPrep

  • Try Free Trial
  • Study Packages
  • Levels I, II & III Lifetime Package
  • Video Lessons
  • Study Notes
  • Practice Questions
  • Levels II & III Lifetime Package
  • About the Exam
  • About your Instructor
  • Part I Study Packages
  • Parts I & II Packages
  • Part I & Part II Lifetime Package
  • Part II Study Packages
  • Exams P & FM Lifetime Package
  • Quantitative Questions
  • Verbal Questions
  • Data Insight Questions
  • Live Tutoring
  • About your Instructors
  • EA Practice Questions
  • Data Sufficiency Questions
  • Integrated Reasoning Questions

Hypothesis Testing

Hypothesis Testing

After completing this reading, you should be able to:

  • Construct an appropriate null hypothesis and alternative hypothesis and distinguish between the two.
  • Construct and apply confidence intervals for one-sided and two-sided hypothesis tests, and interpret the results of hypothesis tests with a specific level of confidence.
  • Differentiate between a one-sided and a two-sided test and identify when to use each test.
  • Explain the difference between Type I and Type II errors and how these relate to the size and power of a test.
  • Understand how a hypothesis test and a confidence interval are related.
  • Explain what the p-value of a hypothesis test measures.
  • Interpret the results of hypothesis tests with a specific level of confidence.
  • Identify the steps to test a hypothesis about the difference between two population means.
  • Explain the problem of multiple testing and how it can bias results.

Hypothesis testing is defined as a process of determining whether a hypothesis is in line with the sample data. Hypothesis testing tries to test whether the observed data of the hypothesis is true. Hypothesis testing starts by stating the null hypothesis and the alternative hypothesis. The null hypothesis is an assumption of the population parameter. On the other hand,  the alternative hypothesis states the parameter values (critical values) at which the null hypothesis is rejected. The critical values are determined by the distribution of the test statistic (when the null hypothesis is true) and the size of the test (which gives the size at which we reject the null hypothesis).

Components of the Hypothesis Testing

The elements of the test hypothesis include:

  • The null hypothesis.
  • The alternative hypothesis.
  • The test statistic.
  • The size of the hypothesis test and errors
  • The critical value.
  • The decision rule.

The Null hypothesis

As stated earlier, the first stage of the hypothesis test is the statement of the null hypothesis. The null hypothesis is the statement concerning the population parameter values. It brings out the notion that “there is nothing about the data.”

The  null hypothesis , denoted as H 0 , represents the current state of knowledge about the population parameter that’s the subject of the test. In other words, it represents the “status quo.” For example, the U.S Food and Drug Administration may walk into a cooking oil manufacturing plant intending to confirm that each 1 kg oil package has, say, 0.15% cholesterol and not more. The inspectors will formulate a hypothesis like:

H 0 : Each 1 kg package has 0.15% cholesterol.

A test would then be carried out to confirm or reject the null hypothesis.

Other typical statements of H 0  include:

$$H_0:\mu={\mu}_0$$

$$H_0:\mu≤{\mu}_0$$

\(μ\) = true population mean and,

\(μ_0\)= the hypothesized population mean.

The Alternative Hypothesis

The  alternative hypothesis , denoted H 1 , is a contradiction of the null hypothesis. The null hypothesis determines the values of the population parameter at which the null hypothesis is rejected. Thus, rejecting the H 0  makes H 1  valid. We accept the alternative hypothesis when the “status quo” is discredited and found to be untrue.

Using our FDA example above, the alternative hypothesis would be:

H 1 : Each 1 kg package does not have 0.15% cholesterol.

The typical statements of H1   include:

$$H_1:\mu \neq {\mu}_0$$

$$H_1:\mu > {\mu}_0$$

Note that we have stated the alternative hypothesis, which contradicted the above statement of the null hypothesis.

The Test Statistic

A test statistic is a standardized value computed from sample information when testing hypotheses. It compares the given data with what we would expect under the null hypothesis. Thus, it is a major determinant when deciding whether to reject H 0 , the null hypothesis.

We use the test statistic to gauge the degree of agreement between sample data and the null hypothesis. Analysts use the following formula when calculating the test statistic.

$$ \text{Test Statistic}= \frac{(\text{Sample Statistic–Hypothesized Value})}{(\text{Standard Error of the Sample Statistic})}$$

The test statistic is a random variable that changes from one sample to another. Test statistics assume a variety of distributions. We shall focus on normally distributed test statistics because it is used hypotheses concerning the means, regression coefficients, and other econometric models.

We shall consider the hypothesis test on the mean. Consider a null hypothesis \(H_0:μ=μ_0\). Assume that the data used is iid, and asymptotic normally distributed as:

$$\sqrt{n} (\hat{\mu}-\mu) \sim N(0, {\sigma}^2)$$

Where \({\sigma}^2\) is the variance of the sequence of the iid random variable used. The asymptotic distribution leads to the test statistic:

$$T=\frac{\hat{\mu}-{\mu}_0}{\sqrt{\frac{\hat{\sigma}^2}{n}}}\sim N(0,1)$$

Note this is consistent with our initial definition of the test statistic.

The following table  gives a brief outline of the various test statistics used regularly, based on the distribution that the data is assumed to follow:

$$\begin{array}{ll} \textbf{Hypothesis Test} & \textbf{Test Statistic}\\ \text{Z-test} & \text{z-statistic} \\ \text{Chi-Square Test} & \text{Chi-Square statistic}\\ \text{t-test} & \text{t-statistic} \\ \text{ANOVA} & \text{F-statistic}\\ \end{array}$$ We can subdivide the set of values that can be taken by the test statistic into two regions: One is called the non-rejection region, which is consistent with H 0  and the rejection region (critical region), which is inconsistent with H 0 . If the test statistic has a value found within the critical region, we reject H 0 .

Just like with any other statistic, the distribution of the test statistic must be specified entirely under H 0  when H 0  is true.

The Size of the Hypothesis Test and the Type I and Type II Errors

While using sample statistics to draw conclusions about the parameters of the population as a whole, there is always the possibility that the sample collected does not accurately represent the population. Consequently, statistical tests carried out using such sample data may yield incorrect results that may lead to erroneous rejection (or lack thereof) of the null hypothesis. We have two types of errors:

Type I Error

Type I error occurs when we reject a true null hypothesis. For example, a type I error would manifest in the form of rejecting H 0  = 0 when it is actually zero.

Type II Error

Type II error occurs when we fail to reject a false null hypothesis. In such a scenario, the test provides insufficient evidence to reject the null hypothesis when it’s false.

The level of significance denoted by α represents the probability of making a type I error, i.e., rejecting the null hypothesis when, in fact, it’s true. α is the direct opposite of β, which is taken to be the probability of making a type II error within the bounds of statistical testing. The ideal but practically impossible statistical test would be one that  simultaneously   minimizes α and β. We use α to determine critical values that subdivide the distribution into the rejection and the non-rejection regions.

The Critical Value and the Decision Rule

The decision to reject or not to reject the null hypothesis is based on the distribution assumed by the test statistic. This means if the variable involved follows a normal distribution, we use the level of significance (α) of the test to come up with critical values that lie along with the standard normal distribution.

The decision rule is a result of combining the critical value (denoted by \(C_α\)), the alternative hypothesis, and the test statistic (T). The decision rule is to whether to reject the null hypothesis in favor of the alternative hypothesis or fail to reject the null hypothesis.

For the t-test, the decision rule is dependent on the alternative hypothesis. When testing the two-side alternative, the decision is to reject the null hypothesis if \(|T|>C_α\). That is, reject the null hypothesis if the absolute value of the test statistic is greater than the critical value. When testing on the one-sided, decision rule, reject the null hypothesis if \(T<C_α\)  when using a one-sided lower alternative and if \(T>C_α\)  when using a one-sided upper alternative. When a null hypothesis is rejected at an α significance level, we say that the result is significant at α significance level.

Note that prior to decision-making, one must decide whether the test should be one-tailed or two-tailed. The following is a brief summary of the decision rules under different scenarios:

Left One-tailed Test

H 1 : parameter < X

Decision rule: Reject H 0  if the test statistic is less than the critical value. Otherwise,  do not reject  H 0.

Right One-tailed Test

H 1 : parameter > X

Decision rule: Reject H 0  if the test statistic is greater than the critical value. Otherwise,  do not reject  H 0.

Two-tailed Test

H 1 : parameter  ≠  X (not equal to X)

Decision rule: Reject H 0  if the test statistic is greater than the upper critical value or less than the lower critical value.

Two-tailed Test

 H 0 : μ < μ 0  vs. H 1 : μ > μ 0.

The second graph represents the rejection region when the alternative is a one-sided upper. The null hypothesis, in this case, is stated as:

H 0 : μ > μ 0  vs. H 1 : μ < μ 0.

Example: Hypothesis Test on the Mean

Consider the returns from a portfolio \(X=(x_1,x_2,\dots, x_n)\) from 1980 through 2020. The approximated mean of the returns is 7.50%, with a standard deviation of 17%. We wish to determine whether the expected value of the return is different from 0 at a 5% significance level.

We start by stating the two-sided hypothesis test:

H 0 : μ =0 vs. H 1 : μ ≠ 0

The test statistic is:

$$T=\frac{\hat{\mu}-{\mu}_0}{\sqrt{\frac{\hat{\sigma}^2}{n}}} \sim N(0,1)$$

In this case, we have,

\(\hat{μ}\)=0.075

\(\hat{\sigma}^2\)=0.17 2

$$T=\frac{0.075-0}{\sqrt{\frac{0.17^2}{40}}} \approx 2.79$$

At the significance level, \(α=5\%\),the critical value is \(±1.96\). Since this is a two-sided test, the rejection regions are ( \(-\infty,-1.96\) ) and (\(1.96, \infty \) ) as shown in the diagram below:

Rejection Regions - Two-Sided Test

The example above is an example of a Z-test (which is mostly emphasized in this chapter and immediately follows from the central limit theorem (CLT)). However, we can use the Student’s t-distribution if the random variables are iid and normally distributed and that the sample size is small (n<30).

In Student’s t-distribution, we used the unbiased estimator of variance. That is:

$$s^2=\frac{\hat{\mu}-{\mu}_0}{\sqrt{\frac{s^2}{n}}}$$

Therefore the test statistic for \(H_0=μ_0\) is given by:

$$T=\frac{\hat{\mu}-{\mu}_0}{\sqrt{\frac{s^2}{n}}} \sim t_{n-1}$$

The Type II Error and the Test Power

The power of a test is the direct opposite of the level of significance. While the level of relevance gives us the probability of rejecting the null hypothesis when it’s, in fact, true, the power of a test gives the probability of correctly discrediting and rejecting the null hypothesis when it is false. In other words, it gives the likelihood of rejecting H 0  when, indeed, it’s false. Denoting the probability of type II error by \(\beta\), the power test is given by:

$$ \text{Power of a Test}=1–\beta $$

The power test measures the likelihood that the false null hypothesis is rejected. It is influenced by the sample size, the length between the hypothesized parameter and the true value, and the size of the test.

Confidence Intervals

A confidence interval can be defined as the range of parameters at which the true parameter can be found at a confidence level. For instance, a 95% confidence interval constitutes the set of parameter values where the null hypothesis cannot be rejected when using a 5% test size. Therefore, a 1-α confidence interval contains values that cannot be disregarded at a test size of α.

It is important to note that the confidence interval depends on the alternative hypothesis statement in the test. Let us start with the two-sided test alternatives.

$$ H_0:μ=0$$

$$H_1:μ≠0$$

Then the \(1-α\) confidence interval is given by:

$$\left[\hat{\mu} -C_{\alpha} \times \frac{\hat {\sigma}}{\sqrt{n}} ,\hat{\mu} + C_{\alpha} \times \frac{\hat {\sigma}}{\sqrt{n}} \right]$$

\(C_α\) is the critical value at \(α\) test size.

Example: Calculating Two-Sided Alternative Confidence Intervals

Consider the returns from a portfolio \(X=(x_1,x_2,…, x_n)\) from 1980 through 2020. The approximated mean of the returns is 7.50%, with a standard deviation of 17%. Calculate the 95% confidence interval for the portfolio return.

The \(1-\alpha\) confidence interval is given by:

$$\begin{align*}&\left[\hat{\mu}-C_{\alpha} \times \frac{\hat {\sigma}}{\sqrt{n}} ,\hat{\mu} + C_{\alpha} \times \frac{\hat {\sigma}}{\sqrt{n}} \right]\\& =\left[0.0750-1.96 \times \frac{0.17}{\sqrt{40}}, 0.0750+1.96 \times \frac{0.17}{\sqrt{40}} \right]\\&=[0.02232,0.1277]\end{align*}$$

Thus, the confidence intervals imply any value of the null between 2.23% and 12.77% cannot be rejected against the alternative.

One-Sided Alternative

For the one-sided alternative, the confidence interval is given by either:

$$\left(-\infty ,\hat{\mu} +C_{\alpha} \times \frac{\hat{\sigma}}{\sqrt{n}} \right )$$

for the lower alternative

$$\left ( \hat{\mu} +C_{\alpha} \times \frac{\hat{\sigma}}{\sqrt{n}},\infty \right )$$

for the upper alternative.

Example: Calculating the One-Sided Alternative Confidence Interval

Assume that we were conducting the following one-sided test:

\(H_0:μ≤0\)

\(H_1:μ>0\)

The 95% confidence interval for the portfolio return is:

$$\begin{align*}&=\left(-\infty ,\hat{\mu} +C_{\alpha} \times \frac{\hat{\sigma}}{\sqrt{n}} \right )\\&=\left(-\infty ,0.0750+1.645\times \frac{0.17}{\sqrt{40}}\right)\\&=(-\infty, 0.1192)\end{align*}$$

On the other hand, if the hypothesis test was:

\(H_0:μ>0\)

\(H_1:μ≤0\)

The 95% confidence interval would be:

$$=\left(-\infty ,\hat{\mu} +C_{\alpha} \times \frac{\hat{\sigma}}{\sqrt{n}} \right )$$

$$=\left(-\infty ,0.0750+1.645\times \frac{0.17}{\sqrt{40}}\right)=(0.1192, \infty)$$

Note that the critical value decreased from 1.96 to 1.645 due to a change in the direction of the change.

The p-Value

When carrying out a statistical test with a fixed value of the significance level (α), we merely compare the observed test statistic with some critical value. For example, we might “reject H 0  using a 5% test” or “reject H 0 at 1% significance level”. The problem with this ‘classical’ approach is that it does not give us details about the  strength of the evidence  against the null hypothesis.

Determination of the  p-value  gives statisticians a more informative approach to hypothesis testing. The p-value is the lowest level at which we can reject H 0 . This means that the strength of the evidence against H 0  increases as the  p-value becomes smaller. The test statistic depends on the alternative.

The p-Value for One-Tailed Test Alternative

For one-tailed tests, the  p-value  is given by the probability that lies below the calculated test statistic for left-tailed tests. Similarly, the likelihood that lies above the test statistic in right-tailed tests gives the  p-value.

Denoting the test statistic by T, the p-value for \(H_1:μ>0\)  is given by:

$$P(Z>|T|)=1-P(Z≤|T|)=1- \Phi (|T|) $$

Conversely , for  \(H_1:μ≤0 \)  the p-value is given by:

$$ P(Z≤|T|)= \Phi (|T|)$$ 

Where z is a standard normal random variable, the absolute value of T (|T|) ensures that the right tail is measured whether T is negative or positive.

The p-Value for Two-Tailed Test Alternative

  If the test is two-tailed, this value is given by the sum of the probabilities in the two tails. We start by determining the probability lying below the negative value of the test statistic. Then, we add this to the probability lying above the positive value of the test statistic. That is the p-value for the two-tailed hypothesis test is given by:

$$2\left[1-\Phi [|T|\right]$$

Example 1: p-Value for One-Sided Alternative

Let θ represent the probability of obtaining a head when a coin is tossed. Suppose we toss the coin 200 times, and heads come up in 85 of the trials. Test the following hypothesis at 5% level of significance.

H 0 : θ = 0.5

H 1 : θ < 0.5

First, not that repeatedly tossing a coin follows a binomial distribution.

Our p-value will be given by P(X < 85) where X  `binomial(200,0.5)  with mean 100(np=200*0.5), assuming H 0  is true.

$$\begin{align*}P\left [ z< \frac{85.5-100}{\sqrt{50}} \right]&=P(Z<-2.05)\\&=1–0.97982=0.02018 \end{align*}$$

Recall that for a binomial distribution, the variance is given by:

$$np(1-p)=200(0.5)(1-0.5)=50$$

(We have applied the Central Limit Theorem by taking the binomial distribution as approx. normal)

Since the probability is less than 0.05, H 0  is extremely unlikely, and we actually have strong evidence against H 0  that favors H 1 . Thus, clearly expressing this result, we could say:

“There is very strong evidence against the hypothesis that the coin is fair. We, therefore, conclude that the coin is biased against heads.”

Remember, failure to reject H 0  does not mean it’s true. It means there’s insufficient evidence to justify rejecting H 0,  given a certain level of significance.

Example 2:  p-Value for Two-Sided Alternative

A CFA candidate conducts a statistical test about the mean value of a random variable X.

H 0 : μ = μ 0  vs. H 1 : μ  ≠  μ 0

She obtains a test statistic of 2.2. Given a 5% significance level, determine and interpret the  p-value

$$ \text{P-value}=2P(Z>2.2)=2[1–P(Z≤2.2)]  =1.39\%×2=2.78\%$$

(We have multiplied by two since this is a two-tailed test)

Example - Two-Sided Test

The p-value (2.78%) is less than the level of significance (5%). Therefore, we have sufficient evidence to reject H 0 . In fact, the evidence is so strong that we would also reject H 0  at significance levels of 4% and 3%. However, at significance levels of 2% or 1%, we would not reject H 0  since the  p-value  surpasses these values.

Hypothesis about the Difference between Two Population Means.

It’s common for analysts to be interested in establishing whether there exists a significant difference between the means of two different populations. For instance, they might want to know whether the average returns for two subsidiaries of a given company exhibit  significant  differences.

Now, consider a bivariate random variable:

$$W_i=[X_i,Y_i]$$

Assume that the components \(X_i\) and \(Y_i\)are both iid and are correlated. That is: \(\text{Corr} (X_i,Y_i )≠0\)

Now, suppose that we want to test the hypothesis that:

$$H_0:μ_X=μ_Y$$

$$H_1:μ_X≠μ_Y$$

In other words, we want to test whether the constituent random variables have equal means. Note that the hypothesis statement above can be written as:

$$H_0:μ_X-μ_Y=0$$

$$H_1:μ_X-μ_Y≠0$$

To execute this test, consider the variable:

$$Z_i=X_i-Y_i$$

Therefore, considering the above random variable, if the null hypothesis is correct then,

$$E(Z_i)=E(X_i)-E(Y_i)=μ_X-μ_Y=0$$

Intuitively, this can be considered as a standard hypothesis test of

H 0 : μ Z =0 vs. H 1 : μ Z  ≠ 0.

The tests statistic is given by:

$$T=\frac{\hat{\mu}_z}{\sqrt{\frac{\hat{\sigma}^2_z}{n}}} \sim N(0,1)$$

Note that the test statistic formula accounts for the correction between \(X_i \) and \(Y_i\). It is easy to see that:

$$V(Z_i)=V(X_i )+V(Y_i)-2COV(X_i, Y_i)$$

Which can be denoted as:

$$\hat{\sigma}^2_z =\hat{\sigma}^2_X +\hat{\sigma}^2_Y – 2{\sigma}_{XY}$$

$$ \hat{\mu}_z ={\mu}_X-{\mu}_Y $$

And thus the test statistic formula can be written as:

$$T=\frac{{\mu}_X -{\mu}_Y}{\sqrt{\frac{\hat{\sigma}^2_X +\hat{\sigma}^2_Y – 2{\sigma}_{XY}}{n}}}$$

This formula indicates that correlation plays a crucial role in determining the magnitude of the test statistic.

Another special case of the test statistic is when \(X_i\), and \(Y_i\) are iid and independent. The test statistic is given by:

$$T=\frac{{\mu}_X -{\mu}_Y}{\sqrt{\frac{\hat{\sigma}^2_X}{n_X}+\frac{\hat{\sigma}^2_Y}{n_Y}}}$$

Where \(n_X\)  and \(n_Y\)  are the sample sizes of \(X_i\), and \(Y_i\) respectively.

Example: Hypothesis Test on Two Means

An investment analyst wants to test whether there is a significant difference between the means of the two portfolios at a 95% level. The first portfolio X consists of 30 government-issued bonds and has a mean of 10% and a standard deviation of 2%. The second portfolio Y consists of 30 private bonds with a mean of 14% and a standard deviation of 3%. The correlation between the two portfolios is 0.7. Calculate the null hypothesis and state whether the null hypothesis is rejected or otherwise.

The hypothesis statement is given by:

H 0 : μ X – μ Y =0 vs. H 1 : μ X – μ Y ≠ 0.

Note that this is a two-tailed test. At 95% level, the test size is α=5% and thus the critical value \(C_α=±1.96\). 

Recall that:

$$Cov(X, Y)=σ_{XY}=ρ_{XY} σ_X σ_Y$$

Where ρ_XY  is the correlation coefficient between X and Y.

Now the test statistic is given by:

$$T=\frac{{\mu}_X -{\mu}_Y}{\sqrt{\frac{\hat{\sigma}^2_X +\hat{\sigma}^2_Y – 2{\sigma}_{XY}}{n}}}=\frac{{\mu}_X -{\mu}_Y}{\sqrt{\frac{\hat{\sigma}^2_X +\hat{\sigma}^2_Y – 2{\rho}_{XY} {\sigma}_X {\sigma}_Y}{n}}}$$

$$=\frac{0.10-0.14}{\sqrt{\frac{0.02^2 +0.03^2-2\times 0.7 \times 0.02 \times 0.03}{30}}}=-10.215$$

The test statistic is far much less than -1.96. Therefore the null hypothesis is rejected at a 95% level.

The Problem of Multiple Testing

Multiple testing occurs when multiple multiple hypothesis tests are conducted on the same data set. The reuse of data results in spurious results and unreliable conclusions that do not hold up to scrutiny. The fundamental problem with multiple testing is that the test size (i.e., the probability that a true null is rejected) is only applicable for a single test. However, repeated testing creates test sizes that are much larger than the assumed size of alpha and therefore increases the probability of a Type I error.

Some control methods have been developed to combat multiple testing. These include Bonferroni correction, the False Discovery Rate (FDR), and Familywise Error Rate (FWER).

Practice Question An experiment was done to find out the number of hours that candidates spend preparing for the FRM part 1 exam. For a sample of 10 students , the average study time was found to be 312.7 hours, with a standard deviation of 7.2 hours. What is the 95% confidence interval for the mean study time of all candidates? A. [307.5, 317.9] B. [310, 317] C. [300, 317] D. [307.5, 312.2] The correct answer is A. To calculate the 95% confidence interval for the mean study time of all candidates, we can use the formula for the confidence interval when the population variance is unknown: \[\text{Confidence Interval} = \bar{X} \pm t_{1-\frac{\alpha}{2}} \times \frac{s}{\sqrt{n}}\] Where: \(\bar{X}\) is the sample mean \(t_{1-\frac{\alpha}{2}}\) is the t-score corresponding to the desired confidence level and degrees of freedom \(s\) is the sample standard deviation \(n\) is the sample size In this case: \(\bar{X} = 312.7\) hours (the average study time) \(s = 7.2\) hours (the standard deviation of study time) \(n = 10\) students (the sample size) To find the t-score (\(t_{1-\frac{\alpha}{2}}\)), we look at the t-table for the 95% confidence level (which corresponds to \(\alpha = 0.05\)) and 9 degrees of freedom (\(n – 1 = 10 – 1 = 9\)). The t-score is 2.262. Now, we can plug these values into the confidence interval formula: \[\text{Confidence Interval} = 312.7 \pm 2.262 \times \frac{7.2}{\sqrt{10}}\] Calculating the margin of error: \[\text{Margin of Error} = 2.262 \times \frac{7.2}{\sqrt{10}} \approx 5.2\] So the confidence interval is: \[\text{Confidence Interval} = 312.7 \pm 5.2 = [307.5, 317.9]\] Therefore, the 95% confidence interval for the mean study time of all candidates is [307.5, 317.9] hours.

Offered by AnalystPrep

null hypothesis confidence interval

Approaches to Asset Allocation

Introduction to derivatives.

After completing this reading, you should be able to: Define derivatives, describe the... Read More

Deciphering the Liquidity and Credit C ...

After completing this reading, you should be able to: Describe the key factors... Read More

Properties of Options

After completing this reading, you should be able to: Identify the six factors... Read More

Pricing Financial Forwards and Futures

After completing this reading, you should be able to: Define and describe financial... Read More

Leave a Comment Cancel reply

You must be logged in to post a comment.

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

Statistics By Jim

Making statistics intuitive

Null Hypothesis: Definition, Rejecting & Examples

By Jim Frost 6 Comments

What is a Null Hypothesis?

The null hypothesis in statistics states that there is no difference between groups or no relationship between variables. It is one of two mutually exclusive hypotheses about a population in a hypothesis test.

Photograph of Rodin's statue, The Thinker who is pondering the null hypothesis.

  • Null Hypothesis H 0 : No effect exists in the population.
  • Alternative Hypothesis H A : The effect exists in the population.

In every study or experiment, researchers assess an effect or relationship. This effect can be the effectiveness of a new drug, building material, or other intervention that has benefits. There is a benefit or connection that the researchers hope to identify. Unfortunately, no effect may exist. In statistics, we call this lack of an effect the null hypothesis. Researchers assume that this notion of no effect is correct until they have enough evidence to suggest otherwise, similar to how a trial presumes innocence.

In this context, the analysts don’t necessarily believe the null hypothesis is correct. In fact, they typically want to reject it because that leads to more exciting finds about an effect or relationship. The new vaccine works!

You can think of it as the default theory that requires sufficiently strong evidence to reject. Like a prosecutor, researchers must collect sufficient evidence to overturn the presumption of no effect. Investigators must work hard to set up a study and a data collection system to obtain evidence that can reject the null hypothesis.

Related post : What is an Effect in Statistics?

Null Hypothesis Examples

Null hypotheses start as research questions that the investigator rephrases as a statement indicating there is no effect or relationship.

Does the vaccine prevent infections? The vaccine does not affect the infection rate.
Does the new additive increase product strength? The additive does not affect mean product strength.
Does the exercise intervention increase bone mineral density? The intervention does not affect bone mineral density.
As screen time increases, does test performance decrease? There is no relationship between screen time and test performance.

After reading these examples, you might think they’re a bit boring and pointless. However, the key is to remember that the null hypothesis defines the condition that the researchers need to discredit before suggesting an effect exists.

Let’s see how you reject the null hypothesis and get to those more exciting findings!

When to Reject the Null Hypothesis

So, you want to reject the null hypothesis, but how and when can you do that? To start, you’ll need to perform a statistical test on your data. The following is an overview of performing a study that uses a hypothesis test.

The first step is to devise a research question and the appropriate null hypothesis. After that, the investigators need to formulate an experimental design and data collection procedures that will allow them to gather data that can answer the research question. Then they collect the data. For more information about designing a scientific study that uses statistics, read my post 5 Steps for Conducting Studies with Statistics .

After data collection is complete, statistics and hypothesis testing enter the picture. Hypothesis testing takes your sample data and evaluates how consistent they are with the null hypothesis. The p-value is a crucial part of the statistical results because it quantifies how strongly the sample data contradict the null hypothesis.

When the sample data provide sufficient evidence, you can reject the null hypothesis. In a hypothesis test, this process involves comparing the p-value to your significance level .

Rejecting the Null Hypothesis

Reject the null hypothesis when the p-value is less than or equal to your significance level. Your sample data favor the alternative hypothesis, which suggests that the effect exists in the population. For a mnemonic device, remember—when the p-value is low, the null must go!

When you can reject the null hypothesis, your results are statistically significant. Learn more about Statistical Significance: Definition & Meaning .

Failing to Reject the Null Hypothesis

Conversely, when the p-value is greater than your significance level, you fail to reject the null hypothesis. The sample data provides insufficient data to conclude that the effect exists in the population. When the p-value is high, the null must fly!

Note that failing to reject the null is not the same as proving it. For more information about the difference, read my post about Failing to Reject the Null .

That’s a very general look at the process. But I hope you can see how the path to more exciting findings depends on being able to rule out the less exciting null hypothesis that states there’s nothing to see here!

Let’s move on to learning how to write the null hypothesis for different types of effects, relationships, and tests.

Related posts : How Hypothesis Tests Work and Interpreting P-values

How to Write a Null Hypothesis

The null hypothesis varies by the type of statistic and hypothesis test. Remember that inferential statistics use samples to draw conclusions about populations. Consequently, when you write a null hypothesis, it must make a claim about the relevant population parameter . Further, that claim usually indicates that the effect does not exist in the population. Below are typical examples of writing a null hypothesis for various parameters and hypothesis tests.

Related posts : Descriptive vs. Inferential Statistics and Populations, Parameters, and Samples in Inferential Statistics

Group Means

T-tests and ANOVA assess the differences between group means. For these tests, the null hypothesis states that there is no difference between group means in the population. In other words, the experimental conditions that define the groups do not affect the mean outcome. Mu (µ) is the population parameter for the mean, and you’ll need to include it in the statement for this type of study.

For example, an experiment compares the mean bone density changes for a new osteoporosis medication. The control group does not receive the medicine, while the treatment group does. The null states that the mean bone density changes for the control and treatment groups are equal.

  • Null Hypothesis H 0 : Group means are equal in the population: µ 1 = µ 2 , or µ 1 – µ 2 = 0
  • Alternative Hypothesis H A : Group means are not equal in the population: µ 1 ≠ µ 2 , or µ 1 – µ 2 ≠ 0.

Group Proportions

Proportions tests assess the differences between group proportions. For these tests, the null hypothesis states that there is no difference between group proportions. Again, the experimental conditions did not affect the proportion of events in the groups. P is the population proportion parameter that you’ll need to include.

For example, a vaccine experiment compares the infection rate in the treatment group to the control group. The treatment group receives the vaccine, while the control group does not. The null states that the infection rates for the control and treatment groups are equal.

  • Null Hypothesis H 0 : Group proportions are equal in the population: p 1 = p 2 .
  • Alternative Hypothesis H A : Group proportions are not equal in the population: p 1 ≠ p 2 .

Correlation and Regression Coefficients

Some studies assess the relationship between two continuous variables rather than differences between groups.

In these studies, analysts often use either correlation or regression analysis . For these tests, the null states that there is no relationship between the variables. Specifically, it says that the correlation or regression coefficient is zero. As one variable increases, there is no tendency for the other variable to increase or decrease. Rho (ρ) is the population correlation parameter and beta (β) is the regression coefficient parameter.

For example, a study assesses the relationship between screen time and test performance. The null states that there is no correlation between this pair of variables. As screen time increases, test performance does not tend to increase or decrease.

  • Null Hypothesis H 0 : The correlation in the population is zero: ρ = 0.
  • Alternative Hypothesis H A : The correlation in the population is not zero: ρ ≠ 0.

For all these cases, the analysts define the hypotheses before the study. After collecting the data, they perform a hypothesis test to determine whether they can reject the null hypothesis.

The preceding examples are all for two-tailed hypothesis tests. To learn about one-tailed tests and how to write a null hypothesis for them, read my post One-Tailed vs. Two-Tailed Tests .

Related post : Understanding Correlation

Neyman, J; Pearson, E. S. (January 1, 1933).  On the Problem of the most Efficient Tests of Statistical Hypotheses .  Philosophical Transactions of the Royal Society A .  231  (694–706): 289–337.

Share this:

null hypothesis confidence interval

Reader Interactions

' src=

January 11, 2024 at 2:57 pm

Thanks for the reply.

January 10, 2024 at 1:23 pm

Hi Jim, In your comment you state that equivalence test null and alternate hypotheses are reversed. For hypothesis tests of data fits to a probability distribution, the null hypothesis is that the probability distribution fits the data. Is this correct?

' src=

January 10, 2024 at 2:15 pm

Those two separate things, equivalence testing and normality tests. But, yes, you’re correct for both.

Hypotheses are switched for equivalence testing. You need to “work” (i.e., collect a large sample of good quality data) to be able to reject the null that the groups are different to be able to conclude they’re the same.

With typical hypothesis tests, if you have low quality data and a low sample size, you’ll fail to reject the null that they’re the same, concluding they’re equivalent. But that’s more a statement about the low quality and small sample size than anything to do with the groups being equal.

So, equivalence testing make you work to obtain a finding that the groups are the same (at least within some amount you define as a trivial difference).

For normality testing, and other distribution tests, the null states that the data follow the distribution (normal or whatever). If you reject the null, you have sufficient evidence to conclude that your sample data don’t follow the probability distribution. That’s a rare case where you hope to fail to reject the null. And it suffers from the problem I describe above where you might fail to reject the null simply because you have a small sample size. In that case, you’d conclude the data follow the probability distribution but it’s more that you don’t have enough data for the test to register the deviation. In this scenario, if you had a larger sample size, you’d reject the null and conclude it doesn’t follow that distribution.

I don’t know of any equivalence testing type approach for distribution fit tests where you’d need to work to show the data follow a distribution, although I haven’t looked for one either!

' src=

February 20, 2022 at 9:26 pm

Is a null hypothesis regularly (always) stated in the negative? “there is no” or “does not”

February 23, 2022 at 9:21 pm

Typically, the null hypothesis includes an equal sign. The null hypothesis states that the population parameter equals a particular value. That value is usually one that represents no effect. In the case of a one-sided hypothesis test, the null still contains an equal sign but it’s “greater than or equal to” or “less than or equal to.” If you wanted to translate the null hypothesis from its native mathematical expression, you could use the expression “there is no effect.” But the mathematical form more specifically states what it’s testing.

It’s the alternative hypothesis that typically contains does not equal.

There are some exceptions. For example, in an equivalence test where the researchers want to show that two things are equal, the null hypothesis states that they’re not equal.

In short, the null hypothesis states the condition that the researchers hope to reject. They need to work hard to set up an experiment and data collection that’ll gather enough evidence to be able to reject the null condition.

' src=

February 15, 2022 at 9:32 am

Dear sir I always read your notes on Research methods.. Kindly tell is there any available Book on all these..wonderfull Urgent

Comments and Questions Cancel reply

Confidence Intervals

Hypothesis testing is the approach to statistical inference that we use when we have two competing theories that we are trying to choose between. A second approach to statistical inference is confidence intervals, which allow us to present a range of reasonable values for our unknown population parameter. The range of reasonable values allows us to understand the corresponding population better without requiring any ideas to be fully specified.

General Motivation and Framework

We have access to our sample, but we would really like to make a statement about the corresponding population. For example, we can calculate that the median price per night for a Chicago Airbnb was \$126 for a sample. What we really want to know, though, is what the median price per night for a Chicago Airbnb is for the entire population of Airbnbs, so that we can make an appropriate statement for the population.

How can we extend our knowledge from the sample to the population? We can use confidence intervals to help us generate a range of reasonable values for our unknown parameter. This will help us to make reasonable conclusions that should extend to the population appropriately.

To do so, we will combine our knowledge of sampling distributions with our specific sample value. This has many similar flavors to hypothesis testing but is approaching the problem through a different framework. Below, we'll walk through an example followed by the process to generate a confidence interval.

Confidence Interval Example

Like mentioned above, the median price per night for a Chicago Airbnb was $126 in our sample. Can we generate a sampling distribution for the possible values that the median price per night could take from repeated random samples?

We will use the resampling approach to generating a sampling distribution as described previously.

null hypothesis confidence interval

Histogram of the sampling distribution for the median price of a Chicago Airbnb.

We've now generated our sampling distribution for sample median prices of Airbnbs in Chicago. Now, suppose that I want to create a range of reasonable values for the population median prices with 90% confidence (we'll define what 90% confidence means soon). To do so, I'll find the middle 90% of this distribution by calculating the 5th percentile and the 95th percentile.

For our simulated sampling distribution, the middle 90% are between \$120 and \$132 per night for a Chicago Airbnb.

At last, we'll make a jump from making statements about samples to making statements about populations. We could say that a range of reasonable values for the population median price per night of a Chicago Airbnb is between \$120 and \$132 per night.

Confidence Interval Steps

To generate a confidence interval, we follow the same set of steps. We do apply some steps differently depending on our specific parameter of interest.

To generate a confidence interval, we should:

  • Identify and define the parameter of interest
  • Determine the confidence level
  • Generate or use theory to specify the sampling distribution and check conditions
  • Calculate the middle region of your sampling distribution, according to your confidence level
  • Write a conclusion in the context of the problem.

Identify Parameter of Interest

We discussed identifying and defining the parameter of interest when we first described hypothesis testing. This is repeated for confidence intervals.

In this example, our population of interest is all Chicago Airbnbs. We likely would want to specify a time frame as well, and since we are using March 2023 data, we may specify that this is for all Chicago Airbnbs in March 2023.

Our parameter of interest (the summary measure) is the median. We may define the parameter of interest as $M$, the population median price per night for a Chicago Airbnb.

Determine the Confidence Level

The confidence level is analogous to the significance level. We'll provide a more exact definition and interpretation of the confidence level shortly. Confidence levels should be greater than 0% and less than 100%.

Confidence levels do not depend on the data and should be selected before observing the data. The confidence level is generally chosen based on the stakeholders and their requirements for the confidence in results. More confidence in the results are associated with higher confidence levels.

Common confidence levels include 90%, 95%, 98%, and 99%.

Determine the Sampling Distribution for the Sample Statistic

We again will use the sampling distribution of the sample statistic as the basis for our confidence interval calculation. To do so, we can follow the same process outlined for hypothesis testing. Recall, that we chose between a simulation-based resampling approach or a theory-based approach using the Central Limit Theorem to define the sampling distribution.

The biggest distinction between generating sampling distributions for confidence intervals compared to hypothesis testing is that we don't need to make any adjustments to our sampling distribution so that it is consistent with the null hypothesis. That is, recall that we wanted to adopt the skeptic's claim in hypothesis testing. When we were generating a sampling distribution, we would make any modifications necessary so that the sampling distribution fulfilled the condition of the null hypothesis. This distinction should be considered in two ways:

  • when generating the sampling distribution
  • when checking any necessary conditions

For example, if we were performing hypothesis testing with a simulation-based approach, we would need to first adjust the data so that the sample median was equal to the null value. However, without that condition for confidence intervals, we would use the data exactly as it is in the sample.

Similarly, some conditions for sampling distributions use information about the parameter of interest. For example, the theory-based approach with proportions requires that $n \times p$ and $n \times (1-p)$ are both at least 10. When we have a hypothesis, we should plug in the null value from the null hypothesis into these checks. With confidence intervals, if we don't have any requirements for the parameter, we can use our best estimate for $p$, which is often $\hat{p}$ when checking the conditions.

Again, the simulation-based approach requires the least number of assumptions. For our example, it is the only option for estimating the sampling distribution, since we haven't introduced theory that relates to the sampling distribution for a sample median.

Calculate the Confidence Interval

After we have determined the sampling distribution, we want to actually calculate the confidence interval, which is the range of reasonable values for our parameter of interest.

We want to find the central part of the sampling distribution that corresponds to our confidence level to generate the confidence interval, regardless of the approach for generating the sampling distribution. That is, if we want a 95% confidence interval, we will want to find the 2.5th percentile and the 97.5th percentile of the sampling distribution, so that the middle 95% is contained within those two values. In general, if we say that our confidence level is represented as CL%, then we want the (100-CL)/2 and (100+CL)/2 percentiles. We can find these percentiles both for a simulated sampling distribution or for a well-defined distribution, as long as we provide Python with the appropriate information.

This might seem counterintuitive, as we are using information about our sample to generate a guess about our population. To understand this, let's start by saying that this range would be a range of typical values for a sample statistic as calculated from our available data. Then, we're going to switch the order of the statement. This indicates that a sample statistic like the one we found would be reasonable if our parameter were anywhere in that range instead. Therefore, we'll say that the confidence interval that we calculated represents a range of reasonable values for the parameter.

Write a Conclusion in the Context of the Problem

Finally, we've generated our confidence interval and want to communicate our results to other stakeholders. What exactly does the confidence interval mean?

Informally, we might say something like: it is reasonable to claim that the population median price for a Chicago Airbnb is between \$120 and \$136 per night, with 90% confidence.

The formal interpretation is that we are 90% confident that the true population median price for a Chicago Airbnb falls in the range of \$120 and \$136 per night.

Confidence Interval Widths

Say that a stakeholder is not satisfied with a confidence interval. A common concern is that a confidence interval is too wide; that is, your stakeholder would like a narrower range of reasonable values. What can be changed to satisfy your stakeholder?

The two adjustable factors that affect the width of the confidence interval are the:

  • sample size
  • confidence level

Larger sample sizes result in narrower sampling distributions (recall this feature of the standard error from our sampling distribution module). This will also result in our confidence interval being narrower.

Larger confidence levels require a larger component of the sampling distribution to be included in the confidence interval. This will result in a wider confidence interval.

Therefore, if your stakeholder wants a narrower confidence interval, you could add more observations to your sample size or you could reduce your confidence level. It is also possible to estimate a desired sample size before gathering data that results in a confidence interval with limitations on the width of the confidence interval. We will skip over this calculation for our course, although you may encounter it in a future course.

Confidence Interval Misconceptions and Misinterpretations

We've discussed briefly what a confidence interval means. Equally important is what a confidence interval does not imply.

A confidence interval does not correspond to:

  • the probability that the parameter is in the confidence interval
  • a range of reasonable values for the sample data
  • a range of reasonable values for a sample statistic
  • a range of reasonable values for any future results from another sample

These last three misconceptions stem from misunderstanding that the confidence interval is about the parameter of interest and not about the sample or any of its corresponding characteristics.

For the first statement, consider that the population is already defined, and the corresponding parameter value for the population could then be calculated. It is a specific number, and it doesn't change. For example, it might be 120 or it could be 145. However, since the population is fixed, it is that exact number.

Once the confidence interval is calculated, then the confidence interval is also set and determined. It won't change. In this case, the parameter will either be contained in our confidence interval or it won't be, so the probability associated with the parameter being in the confidence interval is either 0 (the confidence interval isn't correct) or 1 (the confidence interval is correct).

Confidence Level Interpretation

We now understand how to calculate a confidence interval, what the confidence interval indicates, and what it doesn't indicate. However, we need to return to the second step where we set the confidence level for the interval. We know that this will have ramifications for the following steps of generating a confidence interval. But, what does it mean?

The confidence level means:

"If we gathered repeated random samples of the same size and calculated a CL% confidence interval for each, we would expect CL% of the resulting confidence intervals to contain the true parameter of interest."

Generally, this means that we expect CL% of our intervals to be correct. However, as we discussed above, we can't apply this reasoning to one specific interval after it's been calculated. This still does allow for variability and for different confidence intervals being generated from different samples.

Hypothesis Testing Decisions through Confidence Intervals

You may have noticed that many of the steps used for confidence intervals are shared with hypothesis testing. While there are distinctions between the two, we can also use confidence intervals to help us determine the result of a hypothesis test.

Suppose that a friend found it reported that the median price for all Chicago hotels is $160 per night. They suspect that Airbnbs are less expensive per night, and the population median price for Chicago Airbnbs is less expensive.

That is, the parameter of interest would be $M$ the population median price per night for all Chicago Airbnbs in March 2023. We can (and have) found the corresponding sample statistic, $m$ or the median price per night for the Chicago Airbnbs from our sample.

Because we don't have any data to analyze for Chicago hotels, we'll use this number as if it were true and treat this as a test for only one population. Our hypotheses would be:

$H_0: M = 160$

$H_a: M < 160$

What does the data say? If we've already generated a confidence interval, we don't need to repeat many of the steps for hypothesis testing. Instead, we can consider our calculated confidence interval as a range of reasonable values for our parameter. That is, it is reasonable that the population median price per night for all Chicago Airbnbs is between \$120 and \$136. In this case, the null value of 160 is not included in the range of reasonable values. Everything reasonable falls under the alternative hypothesis. We would want to reject the null hypothesis and adopt the alternative hypothesis as a more reasonable claim.

In this case, our confidence interval clearly supports our alternative hypothesis rather than our null hypothesis. However, in order to use confidence intervals to anticipate the decision for a hypothesis test, we need to ensure that we are using comparable confidence and significance levels:

  • for a two-sided alternative hypothesis, use a confidence level of $1-\alpha$
  • for a one-sided alternative hypothesis, use a confidence level of $1-2\times\alpha$
  • Search Search Please fill out this field.
  • How It Works
  • Calculation

The Bottom Line

  • Fundamental Analysis

What Is a Confidence Interval and How Do You Calculate It?

Adam Hayes, Ph.D., CFA, is a financial writer with 15+ years Wall Street experience as a derivatives trader. Besides his extensive derivative trading expertise, Adam is an expert in economics and behavioral finance. Adam received his master's in economics from The New School for Social Research and his Ph.D. from the University of Wisconsin-Madison in sociology. He is a CFA charterholder as well as holding FINRA Series 7, 55 & 63 licenses. He currently researches and teaches economic sociology and the social studies of finance at the Hebrew University in Jerusalem.

null hypothesis confidence interval

Charlene Rhinehart is a CPA , CFE, chair of an Illinois CPA Society committee, and has a degree in accounting and finance from DePaul University.

null hypothesis confidence interval

Investopedia / Julie Bang

A confidence interval refers to the probability that a population parameter will fall between a set of values for a certain proportion of times.

A confidence interval, in statistics, refers to the probability that a  population  parameter will fall between a set of values for a certain proportion of times. Analysts often use confidence intervals that contain either 95% or 99% of expected observations. Thus, if a point estimate is generated from a statistical model of 10.00 with a 95% confidence interval of 9.50 to 10.50, it means one is 95% confident that the true value falls within that range.

Statisticians and other analysts use confidence intervals to understand the statistical significance of their estimations, inferences, or predictions. If a confidence interval contains the value of zero (or some other null hypothesis) , then one cannot satisfactorily claim that a result from data generated by testing or experimentation is to be attributable to a specific cause rather than chance.

Key Takeaways

  • A confidence interval displays the probability that a parameter will fall between a pair of values around the mean.
  • Confidence intervals measure the degree of uncertainty or certainty in a sampling method.
  • They are also used in hypothesis testing and regression analysis.
  • Statisticians often use p-values in conjunction with confidence intervals to gauge statistical significance.
  • They are most often constructed using confidence levels of 95% or 99%.

Understanding Confidence Intervals

Confidence intervals measure the degree of uncertainty or certainty in a sampling method. They can take any number of probability limits, with the most common being a 95% or 99% confidence level. Confidence intervals are conducted using statistical methods, such as a  t-test .

Statisticians use confidence intervals to measure uncertainty in an estimate of a population parameter based on a sample. For example, a researcher selects different samples randomly from the same population and computes a confidence interval for each sample to see how it may represent the true value of the population variable. The resulting datasets are all different; some intervals include the true population parameter and others do not.

A confidence interval is a range of values, bounded above and below the statistic's mean , that likely would contain an unknown population parameter. Confidence level refers to the percentage of probability, or certainty, that the confidence interval would contain the true population parameter when you draw a random sample many times.

Or, in the vernacular, "we are 99% certain (confidence level) that most of these samples (confidence intervals) contain the true population parameter."

The biggest misconception regarding confidence intervals is that they represent the percentage of data from a given sample that falls between the upper and lower bounds. For example, one might erroneously interpret the aforementioned 99% confidence interval of 70-to-78 inches as indicating that 99% of the data in a random sample falls between these numbers.

This is incorrect, though a separate method of statistical analysis exists to make such a determination. Doing so involves identifying the sample's mean and standard deviation and plotting these figures on a bell curve .

Confidence interval and confidence level are interrelated but are not exactly the same.

Calculating Confidence Intervals

Suppose a group of researchers is studying the heights of high school basketball players. The researchers take a random sample from the population and establish a mean height of 74 inches.

The mean of 74 inches is a point estimate of the population mean. A point estimate by itself is of limited usefulness because it does not reveal the uncertainty associated with the estimate; you do not have a good sense of how far away this 74-inch sample mean might be from the population mean. What's missing is the degree of uncertainty in this single sample.

Confidence intervals provide more information than point estimates. By establishing a 95% confidence interval using the sample's mean and standard deviation , and assuming a normal distribution as represented by the bell curve, the researchers arrive at an upper and lower bound that contains the true mean 95% of the time.

Assume the interval is between 72 inches and 76 inches. If the researchers take 100 random samples from the population of high school basketball players as a whole, the mean should fall between 72 and 76 inches in 95 of those samples.

If the researchers want even greater confidence, they can expand the interval to 99% confidence. Doing so invariably creates a broader range, as it makes room for a greater number of sample means. If they establish the 99% confidence interval as being between 70 inches and 78 inches, they can expect 99 of 100 samples evaluated to contain a mean value between these numbers.

 A 90% confidence level, on the other hand, implies that you would expect 90% of the interval estimates to include the population parameter, and so forth.

What Does a Confidence Interval Reveal?

A confidence interval is a range of values, bounded above and below the statistic's mean, that likely would contain an unknown population parameter. Confidence level refers to the percentage of probability, or certainty, that the confidence interval would contain the true population parameter when you draw a random sample many times.

Why Are Confidence Intervals Used?

Statisticians use confidence intervals to measure uncertainty in a sample variable. For example, a researcher selects different samples randomly from the same population and computes a confidence interval for each sample to see how it may represent the true value of the population variable. The resulting datasets are all different where some intervals include the true population parameter and others do not.

What Is a Common Misconception About Confidence Intervals?

The biggest misconception regarding confidence intervals is that they represent the percentage of data from a given sample that falls between the upper and lower bounds. In other words, it would be incorrect to assume that a 99% confidence interval means that 99% of the data in a random sample falls between these bounds. What it actually means is that one can be 99% certain that the range will contain the population mean.

What Is a T-Test?

Confidence intervals are conducted using statistical methods, such as a t-test. A t-test is a type of inferential statistic used to determine if there is a significant difference between the means of two groups, which may be related to certain features. Calculating a t-test requires three key data values. They include the difference between the mean values from each data set (called the mean difference), the standard deviation of each group, and the number of data values of each group.

How Do You Interpret P-Values and Confidence Intervals?

A p-value is a statistical measurement used to validate a hypothesis against observed data that measures the probability of obtaining the observed results, assuming that the null hypothesis is true. In general, a p-value less than 0.05 is considered to be statistically significant, in which case the null hypothesis should be rejected. This can somewhat correspond to the probability that the null hypothesis value (which is often zero) is contained within a 95% confidence interval.

Confidence intervals allow analysts to understand the likelihood that the results from statistical analyses are real or due to chance. When trying to make inferences or predictions based on a sample of data, there will be some uncertainty as to whether the results of such an analysis actually correspond with the real-world population being studied. The confidence interval depicts the likely range within which the true value should fall.

null hypothesis confidence interval

  • Terms of Service
  • Editorial Policy
  • Privacy Policy

What is The Null Hypothesis & When Do You Reject The Null Hypothesis

Julia Simkus

Editor at Simply Psychology

BA (Hons) Psychology, Princeton University

Julia Simkus is a graduate of Princeton University with a Bachelor of Arts in Psychology. She is currently studying for a Master's Degree in Counseling for Mental Health and Wellness in September 2023. Julia's research has been published in peer reviewed journals.

Learn about our Editorial Process

Saul McLeod, PhD

Editor-in-Chief for Simply Psychology

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

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

Olivia Guy-Evans, MSc

Associate Editor for Simply Psychology

BSc (Hons) Psychology, MSc Psychology of Education

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

On This Page:

A null hypothesis is a statistical concept suggesting no significant difference or relationship between measured variables. It’s the default assumption unless empirical evidence proves otherwise.

The null hypothesis states no relationship exists between the two variables being studied (i.e., one variable does not affect the other).

The null hypothesis is the statement that a researcher or an investigator wants to disprove.

Testing the null hypothesis can tell you whether your results are due to the effects of manipulating ​ the dependent variable or due to random chance. 

How to Write a Null Hypothesis

Null hypotheses (H0) start as research questions that the investigator rephrases as statements indicating no effect or relationship between the independent and dependent variables.

It is a default position that your research aims to challenge or confirm.

For example, if studying the impact of exercise on weight loss, your null hypothesis might be:

There is no significant difference in weight loss between individuals who exercise daily and those who do not.

Examples of Null Hypotheses

Research QuestionNull Hypothesis
Do teenagers use cell phones more than adults?Teenagers and adults use cell phones the same amount.
Do tomato plants exhibit a higher rate of growth when planted in compost rather than in soil?Tomato plants show no difference in growth rates when planted in compost rather than soil.
Does daily meditation decrease the incidence of depression?Daily meditation does not decrease the incidence of depression.
Does daily exercise increase test performance?There is no relationship between daily exercise time and test performance.
Does the new vaccine prevent infections?The vaccine does not affect the infection rate.
Does flossing your teeth affect the number of cavities?Flossing your teeth has no effect on the number of cavities.

When Do We Reject The Null Hypothesis? 

We reject the null hypothesis when the data provide strong enough evidence to conclude that it is likely incorrect. This often occurs when the p-value (probability of observing the data given the null hypothesis is true) is below a predetermined significance level.

If the collected data does not meet the expectation of the null hypothesis, a researcher can conclude that the data lacks sufficient evidence to back up the null hypothesis, and thus the null hypothesis is rejected. 

Rejecting the null hypothesis means that a relationship does exist between a set of variables and the effect is statistically significant ( p > 0.05).

If the data collected from the random sample is not statistically significance , then the null hypothesis will be accepted, and the researchers can conclude that there is no relationship between the variables. 

You need to perform a statistical test on your data in order to evaluate how consistent it is with the null hypothesis. A p-value is one statistical measurement used to validate a hypothesis against observed data.

Calculating the p-value is a critical part of null-hypothesis significance testing because it quantifies how strongly the sample data contradicts the null hypothesis.

The level of statistical significance is often expressed as a  p  -value between 0 and 1. The smaller the p-value, the stronger the evidence that you should reject the null hypothesis.

Probability and statistical significance in ab testing. Statistical significance in a b experiments

Usually, a researcher uses a confidence level of 95% or 99% (p-value of 0.05 or 0.01) as general guidelines to decide if you should reject or keep the null.

When your p-value is less than or equal to your significance level, you reject the null hypothesis.

In other words, smaller p-values are taken as stronger evidence against the null hypothesis. Conversely, when the p-value is greater than your significance level, you fail to reject the null hypothesis.

In this case, the sample data provides insufficient data to conclude that the effect exists in the population.

Because you can never know with complete certainty whether there is an effect in the population, your inferences about a population will sometimes be incorrect.

When you incorrectly reject the null hypothesis, it’s called a type I error. When you incorrectly fail to reject it, it’s called a type II error.

Why Do We Never Accept The Null Hypothesis?

The reason we do not say “accept the null” is because we are always assuming the null hypothesis is true and then conducting a study to see if there is evidence against it. And, even if we don’t find evidence against it, a null hypothesis is not accepted.

A lack of evidence only means that you haven’t proven that something exists. It does not prove that something doesn’t exist. 

It is risky to conclude that the null hypothesis is true merely because we did not find evidence to reject it. It is always possible that researchers elsewhere have disproved the null hypothesis, so we cannot accept it as true, but instead, we state that we failed to reject the null. 

One can either reject the null hypothesis, or fail to reject it, but can never accept it.

Why Do We Use The Null Hypothesis?

We can never prove with 100% certainty that a hypothesis is true; We can only collect evidence that supports a theory. However, testing a hypothesis can set the stage for rejecting or accepting this hypothesis within a certain confidence level.

The null hypothesis is useful because it can tell us whether the results of our study are due to random chance or the manipulation of a variable (with a certain level of confidence).

A null hypothesis is rejected if the measured data is significantly unlikely to have occurred and a null hypothesis is accepted if the observed outcome is consistent with the position held by the null hypothesis.

Rejecting the null hypothesis sets the stage for further experimentation to see if a relationship between two variables exists. 

Hypothesis testing is a critical part of the scientific method as it helps decide whether the results of a research study support a particular theory about a given population. Hypothesis testing is a systematic way of backing up researchers’ predictions with statistical analysis.

It helps provide sufficient statistical evidence that either favors or rejects a certain hypothesis about the population parameter. 

Purpose of a Null Hypothesis 

  • The primary purpose of the null hypothesis is to disprove an assumption. 
  • Whether rejected or accepted, the null hypothesis can help further progress a theory in many scientific cases.
  • A null hypothesis can be used to ascertain how consistent the outcomes of multiple studies are.

Do you always need both a Null Hypothesis and an Alternative Hypothesis?

The null (H0) and alternative (Ha or H1) hypotheses are two competing claims that describe the effect of the independent variable on the dependent variable. They are mutually exclusive, which means that only one of the two hypotheses can be true. 

While the null hypothesis states that there is no effect in the population, an alternative hypothesis states that there is statistical significance between two variables. 

The goal of hypothesis testing is to make inferences about a population based on a sample. In order to undertake hypothesis testing, you must express your research hypothesis as a null and alternative hypothesis. Both hypotheses are required to cover every possible outcome of the study. 

What is the difference between a null hypothesis and an alternative hypothesis?

The alternative hypothesis is the complement to the null hypothesis. The null hypothesis states that there is no effect or no relationship between variables, while the alternative hypothesis claims that there is an effect or relationship in the population.

It is the claim that you expect or hope will be true. The null hypothesis and the alternative hypothesis are always mutually exclusive, meaning that only one can be true at a time.

What are some problems with the null hypothesis?

One major problem with the null hypothesis is that researchers typically will assume that accepting the null is a failure of the experiment. However, accepting or rejecting any hypothesis is a positive result. Even if the null is not refuted, the researchers will still learn something new.

Why can a null hypothesis not be accepted?

We can either reject or fail to reject a null hypothesis, but never accept it. If your test fails to detect an effect, this is not proof that the effect doesn’t exist. It just means that your sample did not have enough evidence to conclude that it exists.

We can’t accept a null hypothesis because a lack of evidence does not prove something that does not exist. Instead, we fail to reject it.

Failing to reject the null indicates that the sample did not provide sufficient enough evidence to conclude that an effect exists.

If the p-value is greater than the significance level, then you fail to reject the null hypothesis.

Is a null hypothesis directional or non-directional?

A hypothesis test can either contain an alternative directional hypothesis or a non-directional alternative hypothesis. A directional hypothesis is one that contains the less than (“<“) or greater than (“>”) sign.

A nondirectional hypothesis contains the not equal sign (“≠”).  However, a null hypothesis is neither directional nor non-directional.

A null hypothesis is a prediction that there will be no change, relationship, or difference between two variables.

The directional hypothesis or nondirectional hypothesis would then be considered alternative hypotheses to the null hypothesis.

Gill, J. (1999). The insignificance of null hypothesis significance testing.  Political research quarterly ,  52 (3), 647-674.

Krueger, J. (2001). Null hypothesis significance testing: On the survival of a flawed method.  American Psychologist ,  56 (1), 16.

Masson, M. E. (2011). A tutorial on a practical Bayesian alternative to null-hypothesis significance testing.  Behavior research methods ,  43 , 679-690.

Nickerson, R. S. (2000). Null hypothesis significance testing: a review of an old and continuing controversy.  Psychological methods ,  5 (2), 241.

Rozeboom, W. W. (1960). The fallacy of the null-hypothesis significance test.  Psychological bulletin ,  57 (5), 416.

Print Friendly, PDF & Email

Replicability-Index

Improving the replicability of empirical research, null-hypothesis testing with confidence intervals.

Statistics is a mess. Statistics education is a mess. Not surprising, the understanding of statistics by applied research workers is a mess. This was less of a problem when there was only one way to conduct statistical analyses. Nobody knew what they were doing, but at least everybody was doing the same thing. Now we have a multiverse of statistical approaches and applied research workers are happy to mix and match statistics to fit their needs. This is making the reporting of results worse and leads to logical contradictions.

For example, the authors of an article in a recent issue of Psychological Science that shall remain anonymous claimed (a) that a Bayesian Hypothesis Test provided evidence for the nil-hypothesis (effect size of zero) and (b) claimed that their preregistered replication study had high statistical power. This makes no sense, because power is defined as the probability of correctly rejecting the null-hypothesis, which assumes an a priori effect size greater than zero. Power is simply not defined when the hypothesis is that the population effect size is zero.

Errors like this are avoidable, if we realize that Neyman introduced confidence intervals to make hypothesis testing easier and more informative. Here is a brief introduction to think clearly about hypothesis testing that should help applied research workers to understand what they are doing.

Effect Size and Sampling Error

The most important information that applied research workers should report are (a) an estimate of the effect size and (b) an estimate of sampling error. Every statistics course should start with introducing these two concepts because all other statistics like p-values or Bayes-Factors or confidence intervals are based on effect size and sampling error. They are also the most useful information for meta-analysis.

Information about effect sizes and sampling error can be in the form of unstandardized values (e.g., 5 cm difference in height with SE = 2 cm) or in standardized form (d = .5, SE = .2). This is not relevant for hypothesis testing and I will use standardized effect sizes for my example.

Specifying the Null-Hypothesis

The null-hypothesis is the hypothesis that a researcher believes to be untrue. It is the hypothesis that they want to reject or NULLify. The biggest mistake in statistics is the assumption that this hypothesis is always that there is no effect (effect size of zero). Cohen (1994) called this hypothesis the nil-hypothesis to distinguish it from other null-hypotheses.

For example, in a directional test that studying harder leads to higher grades, the null-hypothesis specifies all non-positive values (zero and all negative values). When this null-hypothesis is rejected, it automatically implies that the alternative hypothesis is true (given a specific error criterion and a bunch of assumption, not in a mathematically proven sense). Normally, we go through various steps to reject the null-hypothesis, to then affirm the alternative. However, with confidence intervals we can directly affirm the alternative.

Calculating Confidence Intervals

A confidence interval requires three pieces of information.

1. An ESTIMATE of the effect size. This estimate is provided by the mean difference in a sample. For example, the height difference of 5cm or d = .5 are estimates of the population mean difference in height.

2. An Estimate of sampling error. In simple designs, sampling error is a function of sample size, but even then we are making assumptions that can be violated and difficult or impossible to test in small samples In more complex designs, sampling error depends on other statistics that are sample dependent. Thus, sampling error is also just an estimate. The main job of statisticians is to find plausible estimates of sampling error for applied research workers. Applied researchers simply use the information that is provided by statistics programs. In our example, sampling error was estimated to be d = .2.

3. The third piece of information is how confident we want to be in our inferences. All data-based inferences are inductions that can be wrong, but we can specify the probability of being wrong. This quantity is known as the type-I error with the Greek symbol alpha. A common value is alpha = .05. This implies that we have a long-run error rate of no more than 5%. If we obtain 100 confidence intervals, the long-run error rate is limited to no more than 5% false inferences in favor of the alternative hypothesis. With alpha = .05, sampling error has to be multiplied by approximately 2 to compute a confidence interval.

To use our example, with d = .5, and SE = .2, we can create a confidence interval that ranges from d = .5 – .2*2 = .1 to .5 + .2*2 = .9. We can now state that WITHOUT ANY OTHER INFORMATION that may be relevant (e.g., we already know the alternative is true based on a much larger trustworthy prior study and our study is only a classroom demonstration) that the data support our hypothesis that there is a positive effect because the confidence interval fits into the predicted interval; that is values from .1 to .9 fit into the set of values from 0 to infinity.

A more common way to express this finding is to state that the confidence interval does not include the largest value of the null-hypothesis, which is zero. However, this leads to the impression that we tested the nil-hypothesis, and rejected it. But that is not true. We also rejected all the values less than 0. Thus, we did not test or reject the nil-hypothesis. We tested and rejected the null-hypothesis of effect sizes ranging from -infinity to 0. But it is also not necessary to state that we rejected this null-hypothesis because this statement is redundant with the statement we actually want to make. We found evidence for our hypothesis that the effect size is positive (i.e., in the range from 0 to infinity excluding 0).

I hope this example makes it clear how hypothesis testing with confidence intervals works. We first specify a range of values that we think are plausible (e.g., all positive values). We then compute a confidence interval of values that are consistent with our data. We then examine whether the confidence interval falls into the hypothesized range of values. When this is the case, we infer that the data support our hypothesis.

Different Outcomes

When we divide the range of possible effect sizes into two mutually exclusive regions, we can distinguish three possible outcomes.

One possible outcome is that the confidence interval falls into a predicted region. In this case, the data provide support for the prediction.

One possible outcome is that the confidence interval overlaps with the predicted range of values, but also falls outside the range of predicted values. For example, the data could have produced an effect size estimate of d = .1 and an confidence interval ranging from -.3 to .5. In this case, the data are inconclusive. It is possible that the population effect size is, as predicted, positive, or it is negative.

Another possible outcome is that the confidence interval falls entirely outside the predicted range of values (e.g., d = -.5, confidence interval -.9 to -.1). In this case, the data disconfirm the prediction of a positive effect. It follows that it is not even necessary to make a prediction one way or the other. We can simply see whether the confidence interval fits into one or the other region and infer that the population effect size is in the region that contains the confidence interval.

Do We Need A Priori Hypotheses?

Let’s assume that we predicted a positive effect and our hypothesis covers all effect sizes greater than zero and the confidence interval includes values from d = .1 to .9. We said that this finding allows us to accept our hypothesis that the effect size is positive; that is, it is within an interval ranging from 0 to infinity without zero. However, the confidence interval provides a much smaller range of values. A confidence interval ranging from .1 to .9 not only excludes negative values or a value of zero, it also excludes values of 1 or 2. Thus, we are not using all of the information that our data are providing when we simply infer from the data that the effect size is positive, which includes trivial values of 0.0000001 and implausible values of 999.9. The advantage of reporting results with confidence intervals is that we can specify a narrow range of values that are consistent with the data. This is particularly helpful when the confidence interval is close to zero. For example, a confidence interval that ranges from d = 0.001 to d = .801 can be used to claim that the effect size is positive, but it cannot be used to claim that the effect size is theoretically meaningful, unless d = .001 is theoretically meaningful.

Specifying A Minimum Effect Size

To make progress, psychology has to start taking effect sizes more seriously, and this is best achieved by reporting confidence intervals. Confidence intervals ranging from d = .8 to d = 1.2 and ranging from d = .01 to d = .41 are both consistent with the prediction that there is a positive effect, p < .01. However, the two confidence intervals also specify very different ranges of possible effect sizes. Whereas the first confidence interval rejects the hypothesis that effect sizes are small or moderate, the second confidence interval rejects large effect sizes. Traditional hypothesis testing with p-values hides this distinction and makes it look as if these two studies produced identical results. However the lowest value of the first interval (d = .8) is higher than the highest value of the second interval (d = .41), which actually implies that the results are significantly different from each other. Thus, these two studies produced conflicting results when we consider effect sizes, while giving the same answer about the direction of an effect.

If predictions were made in terms of a minimum effect size that is theoretically or practically relevant, the distinction between the two results would also be visible. For example, a standard criterion for a minimum effect size could be a small effect size of d = .2. Using this criterion, the first study confirms predictions (i.e.., the confidence interval from .8 to 1.2 falls into the region from .2 to infinity), but the second study does not, d = .01 to .41 is partially outside the interval from .2 to infinity. In this case, the data are inconclusive.

If the population effect size is zero (e.g., effect of random future events on behavior), confidence intervals will cluster around zero. This makes it hard to fit confidence intervals within a region that is below a minimum effect size (e.g., d = -.2 to d = .2). This is the reason why it is empirically difficult to provide evidence for the absence of an effect. Reducing the minimum effect size makes it even harder and eventually impossible. However, logically there is nothing special about providing evidence for the absence of an effect. We are again dividing the range of plausible effects into two regions: (a) values below the minimum effect size and (b) values above the minimum effect size. We then decide in favor of the interval that fully contains the confidence interval. Of course, we can do this also without an a priori range of effect sizes. For example, if we find a confidence interval ranging from -.15 to +.18, we can infer from this finding that the population effect size is small (less than .2).

But What about Bayesian Statistics?

Bayesian statistics also uses information about effect sizes and sampling error. The main difference is that Bayesians assume that we have prior knowledge that can inform our interpretation of results. For example, if one-hundred studies already tested the same hypothesis, we can use the information of these studies. In this case, it would also be possible to conduct a meta-analysis and to draw inferences on evidence from all 101 studies, rather than just a single study. Bayesians also sometimes incorporate information that is harder to quantify. However, the main logic of hypothesis testing with confidence intervals or Bayesian credibility intervals does not change. Ironically, Bayesians also tend to use alpha = .05 when they report 95% credibility intervals. The only difference is that information that is external to the data (prior distributions) is used, whereas confidence intervals rely exclusively on information from the data.

I hope that this blog post helps researchers to better understand what they are doing. Empirical studies provides estimates of two important statistics, an effect size estimate and an sampling error estimate. This information can be used to create intervals that specify a range of values that are likely to contain the population effect size. Hypotheses testing divides the range of possible values into regions and decides in favor of hypotheses that fully contain the confidence interval. However, hypothesis testing is redundant and less informative because we can simply decide in favor of the values that are inside the confidence interval which is smaller than the range of values specified by a theoretical prediction. The use of confidence intervals makes it possible to identify weak evidence (confidence interval excludes zero, but not very small values that are not theoretically interesting) and also makes it possible to provide evidence for the absence of an effect (confidence interval only includes trivial values).

A common criticism of hypothesis testing is that it is difficult to understand and not intuitive. The use of confidence intervals solves this problem. Seeing whether a small objects fits into a larger object is probably achieved at some early developmental stage in Piaget’s model and most applied research workers should be able to carry out these comparisons. Standardized effect sizes also help with evaluating the size of objects. Thus, confidence intervals provide all of the information that applied research workers need to carry out empirical studies and to draw inferences from these studies. The main statistical challenge is to obtain estimates of sampling error in complex designs; that is the job of statisticians. The main job of empirical research workers is to collect theoretically or practically important data with small sampling error.

Share this:

10 thoughts on “ null-hypothesis testing with confidence intervals ”.

I’m not a researcher, but I appreciate your site.

Great post. I plan to share this with other research friends. The next step is improving statistical training and literacy in the young generation of researchers like myself.

Thank you for the feedback.

“It is the hypothesis that they want to reject or NULLify”. Perhaps just a poor choice of words, but “want to reject” is at the heart of so many of social psych’s problems.

I see what you are saying but the only thing you can do with the nil-hypothesis is to reject it. So the nature of the null-hypothesis is the reason for the problem that everybody wants to reject it.

  • Pingback: Expressing Uncertainty about Analysis Plans with Conservative Confidence Intervals | Replicability-Index

Hi Ulrich, “Power is simply not defined when the hypothesis is that the population effect size is zero.” Maybe I do not understand this sentence correctly, but I think, this is not correct, Power is simply defined as the probability to identify a particular effect different from the effect under study. However, it is not a feature of an underlying distribution, but of a decision-theoretic comparison between two distributions, in the most simple case of two sampling distributions of t (we call it t-test). Mathematically, the form of the central t distribution (aka Student’s t) is not much different from non-central t distributions, except for the non-centrality parameter (which is 0, reflecting the absence of any effect, for the central t distribution). Assume now, we are testing our research hypothesis of, e.g., d>=0.5 as meaningful non-nil null hypothesis (d_0) (since we are hard-boned falsificationists!). The alternative is h1: d<0.5. Then, the probabilty to identify a possibly correct alternative d_1 (which actually is Power, right?) will be higher the larger the difference between d_0 and d_1 is, right? Which means that the Power to identify an alternative correctly is higher for d_1=0 than for larger values of d, e.g. d_1=0.3. Which also means that Power for the correct identification for a true population effect of d=0 does not only exist, but is also computable.

Hi Uwe, I am follow the common definition of power as a conditional probability. “The power of a binary hypothesis test is the probability that the test rejects the null hypothesis (H0) when a specific alternative hypothesis (H1) is true.” https://en.wikipedia.org/wiki/Power_(statistics)

However, I have no problem to use the term also for cases where the null is true and “power” equals alpha.

As long as it is clear what we are talking about, both are important and useful concepts.

But doesn’t my example exactly fit the definition from Wikipedia, provided my null hypothesis (d=0.5, assumed effect) is false and a specific alternative (d=0, no effect) is true?

Maybe I am not understanding your point. If I am testing d .5, power would not be defined if the true population parameter is d = .5, No matter how many subjects I have, I will only get significant results at the rate of alpha.

Leave a Reply Cancel reply

Discover more from replicability-index.

Subscribe now to keep reading and get access to the full archive.

Type your email…

Continue reading

Pardon Our Interruption

As you were browsing something about your browser made us think you were a bot. There are a few reasons this might happen:

  • You've disabled JavaScript in your web browser.
  • You're a power user moving through this website with super-human speed.
  • You've disabled cookies in your web browser.
  • A third-party browser plugin, such as Ghostery or NoScript, is preventing JavaScript from running. Additional information is available in this support article .

To regain access, please make sure that cookies and JavaScript are enabled before reloading the page.

Open Access Article

Title: Alpha, Null Hypothesis Statistical Testing and Confidence Intervals: Where Do We Go From Here?

Authors : Tamela D. Ferguson; Wlliam L. Ferguson

Addresses : N/A ' N/A

Abstract : The choice of alpha = .05 to determine significance in null hypothesis statistical testing (NHST) has become ingrained in management research, though the choice of .05 level appears to have little scientific basis and is likely simple convenience. More appropriate approaches might be to choose an alpha level based on specific research design or stage of research stream development. Over the years, many have criticized NHST use, with some calling for its ban while advocating the use of confidence intervals instead. To this end, this paper presents a historical review, support for and criticisms of alpha and the related NHST, as well as discussion of issues concerning the use of confidence intervals as an alternative to NHST. The development of the organizational configurations-performance stream in strategic research is used to illustrate the critical importance of researchers making appropriate statistical testing choices.

Keywords : Alpha; null hypothesis statistical testing; confidence intervals; configurations-performance stream.

DOI : 10.1504/JBM.2002.141077

Journal of Business and Management, 2002 Vol.8 No.1, pp.87 - 98

Published online: 05 Sep 2024 *

Keep up-to-date

  • Our Newsletter ( subscribe for free )
  • New issue alerts
  • Inderscience is a member of publishing organisations including:

CLOCKSS

IMAGES

  1. The Relationship Between Confidence Intervals and Hypothesis Tests

    null hypothesis confidence interval

  2. Hypothesis Testing

    null hypothesis confidence interval

  3. Understanding Hypothesis Tests: Confidence Intervals and Confidence Levels

    null hypothesis confidence interval

  4. Use of confidence interval in hypothesis testing.

    null hypothesis confidence interval

  5. Hypothesis Testing and Confidence Intervals

    null hypothesis confidence interval

  6. Understanding Hypothesis Tests: Confidence Intervals and Confidence Levels

    null hypothesis confidence interval

VIDEO

  1. Test of Hypothesis using Python

  2. Hypothesis Tests| Some Concepts

  3. 114 Testing of Hypothesis & Confidence Interval: Lecture V

  4. Full Lecture: Confidence Intervals for Standard Deviation and General Hypothesis Testing

  5. LCHL

  6. Hypothesis Testing

COMMENTS

  1. 6.6

    In other words, if the the 95% confidence interval contains the hypothesized parameter, then a hypothesis test at the 0.05 \(\alpha\) level will almost always fail to reject the null hypothesis. If the 95% confidence interval does not contain the hypothesize parameter, then a hypothesis test at the 0.05 \(\alpha\) level will almost always ...

  2. Hypothesis Testing and Confidence Intervals

    Confidence intervals and hypothesis testing are closely related because both methods use the same underlying methodology. Additionally, there is a close connection between significance levels and confidence levels. ... Confidence interval approach: If the null hypothesis mean is more than $63.57 from the sample mean, the interval does not ...

  3. Hypothesis Testing, P Values, Confidence Intervals, and Significance

    Medical providers often rely on evidence-based medicine to guide decision-making in practice. Often a research hypothesis is tested with results provided, typically with p values, confidence intervals, or both. Additionally, statistical or research significance is estimated or determined by the investigators. Unfortunately, healthcare providers may have different comfort levels in interpreting ...

  4. Understanding Confidence Intervals

    To calculate the 95% confidence interval, we can simply plug the values into the formula. For the USA: So for the USA, the lower and upper bounds of the 95% confidence interval are 34.02 and 35.98. For GB: So for the GB, the lower and upper bounds of the 95% confidence interval are 33.04 and 36.96.

  5. Statistical tests, P values, confidence intervals, and power: a guide

    Keywords: Confidence intervals, Hypothesis testing, Null testing, P value, Power, Significance tests, Statistical testing. ... If one 95 % confidence interval includes the null value and another excludes that value, the interval excluding the null is the more precise one. No!

  6. 11.8: Significance Testing and Confidence Intervals

    If the \(95\%\) confidence interval contains zero (more precisely, the parameter value specified in the null hypothesis), then the effect will not be significant at the \(0.05\) level. Looking at non-significant effects in terms of confidence intervals makes clear why the null hypothesis should not be accepted when it is not rejected: Every ...

  7. The Relationship Between Hypothesis Testing and Confidence Intervals

    Hypothesis tests are centered around the null hypothesized parameter and confidence intervals are centered around the estimate of the sample parameter. How can use these two concepts in tandem? If the null hypothesized value is found in our confidence interval, then that would mean we have a bad confidence interval and our p-value would be high.

  8. 8.6 Relationship Between Confidence Intervals and Hypothesis Tests

    We can also use confidence intervals to make conclusions about hypothesis tests: reject the null hypothesis [latex]H_0[/latex] at the significance level [latex]\alpha[/latex] if the corresponding [latex](1 - \alpha) \times 100\%[/latex] confidence interval does not contain the hypothesized value [latex]\mu_0[/latex]. The relationship is ...

  9. Understanding Hypothesis Tests: Confidence Intervals and ...

    You can use either P values or confidence intervals to determine whether your results are statistically significant. If a hypothesis test produces both, these results will agree. The confidence level is equivalent to 1 - the alpha level. So, if your significance level is 0.05, the corresponding confidence level is 95%.

  10. Confidence Intervals: Interpreting, Finding & Formulas

    A confidence interval (CI) is a range of values that is likely to contain the value of an unknown population parameter. These intervals represent a plausible domain for the parameter given the characteristics of your sample data. Confidence intervals are derived from sample statistics and are calculated using a specified confidence level.

  11. The Ultimate Guide to Hypothesis Testing and Confidence Intervals in

    By comparing the statistics calculated here, we can decide whether to reject the null hypothesis or not. Confidence Interval. The confidence interval for proportion follows the same pattern as the statistical inference for mean, which is using the point estimate and margin of errors, except the standard errors are calculated differently here :

  12. Null hypothesis significance testing: a short tutorial

    Keywords: null hypothesis significance testing, tutorial, p-value, reporting, confidence intervals The Null Hypothesis Significance Testing framework NHST is a method of statistical inference by which an experimental factor is tested against a hypothesis of no effect or no relationship based on a given observation.

  13. Confidence Intervals and p-Values

    Therefore, if the null value (RR=1.0 or OR=1.0) is not contained within the 95% confidence interval, then the probability that the null is the true value is less than 5%. Conversely, if the null is contained within the 95% confidence interval, then the null is one of the values that is consistent with the observed data, so the null hypothesis ...

  14. Using a confidence interval to decide whether to reject the null hypothesis

    Using a confidence interval to decide whether to reject the null hypothesis. Suppose that you do a hypothesis test. Remember that the decision to reject the null hypothesis (H 0) or fail to reject it can be based on the p-value and your chosen significance level (also called α). If the p-value is less than or equal to α, you reject H 0; if it ...

  15. Rejecting the Null Hypothesis Using Confidence Intervals

    As a hypothesis test, we could have the alternative hypothesis as H 1 ≠ 0.51. Since the null value of 0.51 lies within the confidence interval, then we would fail to reject the null hypothesis at ɑ = 0.05. On the other hand, if H 1 ≠ 0.61, then since 0.61 is not in the confidence interval we can reject the null hypothesis at ɑ = 0.05.

  16. Hypothesis Testing and Confidence Intervals

    A confidence interval can be defined as the range of parameters at which the true parameter can be found at a confidence level. For instance, a 95% confidence interval constitutes the set of parameter values where the null hypothesis cannot be rejected when using a 5% test size.

  17. Null Hypothesis: Definition, Rejecting & Examples

    When your sample contains sufficient evidence, you can reject the null and conclude that the effect is statistically significant. Statisticians often denote the null hypothesis as H 0 or H A.. Null Hypothesis H 0: No effect exists in the population.; Alternative Hypothesis H A: The effect exists in the population.; In every study or experiment, researchers assess an effect or relationship.

  18. Confidence Intervals

    The biggest distinction between generating sampling distributions for confidence intervals compared to hypothesis testing is that we don't need to make any adjustments to our sampling distribution so that it is consistent with the null hypothesis. That is, recall that we wanted to adopt the skeptic's claim in hypothesis testing. ...

  19. Null hypothesis

    The null hypothesis and the alternative hypothesis are types of conjectures used in statistical tests to make statistical inferences, which are formal methods of reaching conclusions and separating scientific claims from statistical noise.. The statement being tested in a test of statistical significance is called the null hypothesis. The test of significance is designed to assess the strength ...

  20. 12: Confidence Intervals and Hypothesis Tests

    In this chapter, you will learn to construct and interpret confidence intervals. You will also learn a new distribution, the Student's-t, and how it is used with these intervals. ... to reject the null hypothesis. 12.3.1: Null and Alternative Hypotheses; 12.3.2: Outcomes and the Type I and Type II Errors; 12.3.3: Distribution Needed for ...

  21. What Is a Confidence Interval and How Do You Calculate It?

    Investopedia / Julie Bang. A confidence interval, in statistics, refers to the probability that a population parameter will fall between a set of values for a certain proportion of times. Analysts ...

  22. What Is The Null Hypothesis & When To Reject It

    A tutorial on a practical Bayesian alternative to null-hypothesis significance testing. Behavior research methods, 43, 679-690. Nickerson, R. S. (2000). Null hypothesis significance testing: a review of an old and continuing controversy. Psychological methods, 5(2), 241. Rozeboom, W. W. (1960). The fallacy of the null-hypothesis significance test.

  23. Null-Hypothesis Testing with Confidence Intervals

    A more common way to express this finding is to state that the confidence interval does not include the largest value of the null-hypothesis, which is zero. However, this leads to the impression that we tested the nil-hypothesis, and rejected it. But that is not true. We also rejected all the values less than 0.

  24. Statistics Exam: Confidence Intervals, Variance, Hypothesis

    Develop a 95% confidence interval estimate of the mean amount spent per day by a couple couples vacationing at the Resort XYZ. Question-2 (10 points) Part variability is critical in the manufacturing of ball bearings. ... At the ? = .05 level of significance, test the null hypothesis. Question-8 ...

  25. Article: Alpha, Null Hypothesis Statistical Testing and Confidence

    Journal of Business and Management; 2002 Vol.8 No.1; Read the full-text of this article for free. Title: Alpha, Null Hypothesis Statistical Testing and Confidence Intervals: Where Do We Go From Here? Authors: Tamela D. Ferguson; Wlliam L. Ferguson. Addresses: N/A ' N/A. Abstract: The choice of alpha = .05 to determine significance in null hypothesis statistical testing (NHST) has become ...