Have a language expert improve your writing

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

  • Knowledge Base

Methodology

  • Random vs. Systematic Error | Definition & Examples

Random vs. Systematic Error | Definition & Examples

Published on May 7, 2021 by Pritha Bhandari . Revised on June 22, 2023.

In scientific research, measurement error is the difference between an observed value and the true value of something. It’s also called observation error or experimental error.

There are two main types of measurement error:

Random error is a chance difference between the observed and true values of something (e.g., a researcher misreading a weighing scale records an incorrect measurement).

  • Systematic error is a consistent or proportional difference between the observed and true values of something (e.g., a miscalibrated scale consistently registers weights as higher than they actually are).

By recognizing the sources of error, you can reduce their impacts and record accurate and precise measurements. Gone unnoticed, these errors can lead to research biases like omitted variable bias or information bias .

Table of contents

Are random or systematic errors worse, random error, reducing random error, systematic error, reducing systematic error, other interesting articles, frequently asked questions about random and systematic error.

In research, systematic errors are generally a bigger problem than random errors.

Random error isn’t necessarily a mistake, but rather a natural part of measurement. There is always some variability in measurements, even when you measure the same thing repeatedly, because of fluctuations in the environment, the instrument, or your own interpretations.

But variability can be a problem when it affects your ability to draw valid conclusions about relationships between variables . This is more likely to occur as a result of systematic error.

Precision vs accuracy

Random error mainly affects precision , which is how reproducible the same measurement is under equivalent circumstances. In contrast, systematic error affects the accuracy of a measurement, or how close the observed value is to the true value.

Taking measurements is similar to hitting a central target on a dartboard. For accurate measurements, you aim to get your dart (your observations) as close to the target (the true values) as you possibly can. For precise measurements, you aim to get repeated observations as close to each other as possible.

Random error introduces variability between different measurements of the same thing, while systematic error skews your measurement away from the true value in a specific direction.

Precision vs accuracy

When you only have random error, if you measure the same thing multiple times, your measurements will tend to cluster or vary around the true value. Some values will be higher than the true score, while others will be lower. When you average out these measurements, you’ll get very close to the true score.

For this reason, random error isn’t considered a big problem when you’re collecting data from a large sample—the errors in different directions will cancel each other out when you calculate descriptive statistics . But it could affect the precision of your dataset when you have a small sample.

Systematic errors are much more problematic than random errors because they can skew your data to lead you to false conclusions. If you have systematic error, your measurements will be biased away from the true values. Ultimately, you might make a false positive or a false negative conclusion (a Type I or II error ) about the relationship between the variables you’re studying.

Here's why students love Scribbr's proofreading services

Discover proofreading & editing

Random error affects your measurements in unpredictable ways: your measurements are equally likely to be higher or lower than the true values.

In the graph below, the black line represents a perfect match between the true scores and observed scores of a scale. In an ideal world, all of your data would fall on exactly that line. The green dots represent the actual observed scores for each measurement with random error added.

Random error

Random error is referred to as “noise”, because it blurs the true value (or the “signal”) of what’s being measured. Keeping random error low helps you collect precise data.

Sources of random errors

Some common sources of random error include:

  • natural variations in real world or experimental contexts.
  • imprecise or unreliable measurement instruments.
  • individual differences between participants or units.
  • poorly controlled experimental procedures.
Random error source Example
Natural variations in context In an about memory capacity, your participants are scheduled for memory tests at different times of day. However, some participants tend to perform better in the morning while others perform better later in the day, so your measurements do not reflect the true extent of memory capacity for each individual.
Imprecise instrument You measure wrist circumference using a tape measure. But your tape measure is only accurate to the nearest half-centimeter, so you round each measurement up or down when you record data.
Individual differences You ask participants to administer a safe electric shock to themselves and rate their pain level on a 7-point rating scale. Because pain is subjective, it’s hard to reliably measure. Some participants overstate their levels of pain, while others understate their levels of pain.

Random error is almost always present in research, even in highly controlled settings. While you can’t eradicate it completely, you can reduce random error using the following methods.

Take repeated measurements

A simple way to increase precision is by taking repeated measurements and using their average. For example, you might measure the wrist circumference of a participant three times and get slightly different lengths each time. Taking the mean of the three measurements, instead of using just one, brings you much closer to the true value.

Increase your sample size

Large samples have less random error than small samples. That’s because the errors in different directions cancel each other out more efficiently when you have more data points. Collecting data from a large sample increases precision and statistical power .

Control variables

In controlled experiments , you should carefully control any extraneous variables that could impact your measurements. These should be controlled for all participants so that you remove key sources of random error across the board.

Systematic error means that your measurements of the same thing will vary in predictable ways: every measurement will differ from the true measurement in the same direction, and even by the same amount in some cases.

Systematic error is also referred to as bias because your data is skewed in standardized ways that hide the true values. This may lead to inaccurate conclusions.

Types of systematic errors

Offset errors and scale factor errors are two quantifiable types of systematic error.

An offset error occurs when a scale isn’t calibrated to a correct zero point. It’s also called an additive error or a zero-setting error.

A scale factor error is when measurements consistently differ from the true value proportionally (e.g., by 10%). It’s also referred to as a correlational systematic error or a multiplier error.

You can plot offset errors and scale factor errors in graphs to identify their differences. In the graphs below, the black line shows when your observed value is the exact true value, and there is no random error.

The blue line is an offset error: it shifts all of your observed values upwards or downwards by a fixed amount (here, it’s one additional unit).

The purple line is a scale factor error: all of your observed values are multiplied by a factor—all values are shifted in the same direction by the same proportion, but by different absolute amounts.

Systematic error

Sources of systematic errors

The sources of systematic error can range from your research materials to your data collection procedures and to your analysis techniques. This isn’t an exhaustive list of systematic error sources, because they can come from all aspects of research.

Response bias occurs when your research materials (e.g., questionnaires ) prompt participants to answer or act in inauthentic ways through leading questions . For example, social desirability bias can lead participants try to conform to societal norms, even if that’s not how they truly feel.

Your question states: “Experts believe that only systematic actions can reduce the effects of climate change. Do you agree that individual actions are pointless?”

Experimenter drift occurs when observers become fatigued, bored, or less motivated after long periods of data collection or coding, and they slowly depart from using standardized procedures in identifiable ways.

Initially, you code all subtle and obvious behaviors that fit your criteria as cooperative. But after spending days on this task, you only code extremely obviously helpful actions as cooperative.

Sampling bias occurs when some members of a population are more likely to be included in your study than others. It reduces the generalizability of your findings, because your sample isn’t representative of the whole population.

You can reduce systematic errors by implementing these methods in your study.

Triangulation

Triangulation means using multiple techniques to record observations so that you’re not relying on only one instrument or method.

For example, if you’re measuring stress levels, you can use survey responses, physiological recordings, and reaction times as indicators. You can check whether all three of these measurements converge or overlap to make sure that your results don’t depend on the exact instrument used.

Regular calibration

Calibrating an instrument means comparing what the instrument records with the true value of a known, standard quantity. Regularly calibrating your instrument with an accurate reference helps reduce the likelihood of systematic errors affecting your study.

You can also calibrate observers or researchers in terms of how they code or record data. Use standard protocols and routine checks to avoid experimenter drift.

Randomization

Probability sampling methods help ensure that your sample doesn’t systematically differ from the population.

In addition, if you’re doing an experiment, use random assignment to place participants into different treatment conditions. This helps counter bias by balancing participant characteristics across groups.

Wherever possible, you should hide the condition assignment from participants and researchers through masking (blinding) .

Participants’ behaviors or responses can be influenced by experimenter expectancies and demand characteristics in the environment, so controlling these will help you reduce systematic bias.

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

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

Research bias

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

Random and systematic error are two types of measurement error.

Systematic error is a consistent or proportional difference between the observed and true values of something (e.g., a miscalibrated scale consistently records weights as higher than they actually are).

Systematic error is generally a bigger problem in research.

With random error, multiple measurements will tend to cluster around the true value. When you’re collecting data from a large sample , the errors in different directions will cancel each other out.

Systematic errors are much more problematic because they can skew your data away from the true value. This can lead you to false conclusions ( Type I and II errors ) about the relationship between the variables you’re studying.

Random error  is almost always present in scientific studies, even in highly controlled settings. While you can’t eradicate it completely, you can reduce random error by taking repeated measurements, using a large sample, and controlling extraneous variables .

You can avoid systematic error through careful design of your sampling , data collection , and analysis procedures. For example, use triangulation to measure your variables using multiple methods; regularly calibrate instruments or procedures; use random sampling and random assignment ; and apply masking (blinding) where possible.

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.

Bhandari, P. (2023, June 22). Random vs. Systematic Error | Definition & Examples. Scribbr. Retrieved September 16, 2024, from https://www.scribbr.com/methodology/random-vs-systematic-error/

Is this article helpful?

Pritha Bhandari

Pritha Bhandari

Other students also liked, reliability vs. validity in research | difference, types and examples, what is a controlled experiment | definitions & examples, extraneous variables | examples, types & controls, what is your plagiarism score.

Types of Error — Overview & Comparison - Expii

  • WolframAlpha.com
  • WolframCloud.com
  • All Sites & Public Resources...

short definition of experimental error

  • Wolfram|One
  • Mathematica
  • Wolfram|Alpha Notebook Edition
  • Finance Platform
  • System Modeler
  • Wolfram Player
  • Wolfram Engine
  • WolframScript
  • Enterprise Private Cloud
  • Application Server
  • Enterprise Mathematica
  • Wolfram|Alpha Appliance
  • Corporate Consulting
  • Technical Consulting
  • Wolfram|Alpha Business Solutions
  • Data Repository
  • Neural Net Repository
  • Function Repository
  • Wolfram|Alpha Pro
  • Problem Generator
  • Products for Education
  • Wolfram Cloud App
  • Wolfram|Alpha for Mobile
  • Wolfram|Alpha-Powered Apps
  • Paid Project Support
  • Summer Programs
  • All Products & Services »
  • Wolfram Language Revolutionary knowledge-based programming language. Wolfram Cloud Central infrastructure for Wolfram's cloud products & services. Wolfram Science Technology-enabling science of the computational universe. Wolfram Notebooks The preeminent environment for any technical workflows. Wolfram Engine Software engine implementing the Wolfram Language. Wolfram Natural Language Understanding System Knowledge-based broadly deployed natural language. Wolfram Data Framework Semantic framework for real-world data. Wolfram Universal Deployment System Instant deployment across cloud, desktop, mobile, and more. Wolfram Knowledgebase Curated computable knowledge powering Wolfram|Alpha.
  • All Technologies »
  • Aerospace & Defense
  • Chemical Engineering
  • Control Systems
  • Electrical Engineering
  • Image Processing
  • Industrial Engineering
  • Mechanical Engineering
  • Operations Research
  • Actuarial Sciences
  • Bioinformatics
  • Data Science
  • Econometrics
  • Financial Risk Management
  • All Solutions for Education
  • Machine Learning
  • Multiparadigm Data Science
  • High-Performance Computing
  • Quantum Computation Framework
  • Software Development
  • Authoring & Publishing
  • Interface Development
  • Web Development
  • All Solutions »
  • Wolfram Language Documentation
  • Fast Introduction for Programmers
  • Videos & Screencasts
  • Wolfram Language Introductory Book
  • Webinars & Training
  • Support FAQ
  • Wolfram Community
  • Contact Support
  • All Learning & Support »
  • Company Background
  • Wolfram Blog
  • Careers at Wolfram
  • Internships
  • Other Wolfram Language Jobs
  • Wolfram Foundation
  • Computer-Based Math
  • A New Kind of Science
  • Wolfram Technology for Hackathons
  • Student Ambassador Program
  • Wolfram for Startups
  • Demonstrations Project
  • Wolfram Innovator Awards
  • Wolfram + Raspberry Pi
  • All Company »

Chapter 3

Experimental Errors and

Error Analysis

This chapter is largely a tutorial on handling experimental errors of measurement. Much of the material has been extensively tested with science undergraduates at a variety of levels at the University of Toronto.

Whole books can and have been written on this topic but here we distill the topic down to the essentials. Nonetheless, our experience is that for beginners an iterative approach to this material works best. This means that the users first scan the material in this chapter; then try to use the material on their own experiment; then go over the material again; then ...

provides functions to ease the calculations required by propagation of errors, and those functions are introduced in Section 3.3. These error propagation functions are summarized in Section 3.5.

3.1 Introduction

3.1.1 The Purpose of Error Analysis

For students who only attend lectures and read textbooks in the sciences, it is easy to get the incorrect impression that the physical sciences are concerned with manipulating precise and perfect numbers. Lectures and textbooks often contain phrases like:

For an experimental scientist this specification is incomplete. Does it mean that the acceleration is closer to 9.8 than to 9.9 or 9.7? Does it mean that the acceleration is closer to 9.80000 than to 9.80001 or 9.79999? Often the answer depends on the context. If a carpenter says a length is "just 8 inches" that probably means the length is closer to 8 0/16 in. than to 8 1/16 in. or 7 15/16 in. If a machinist says a length is "just 200 millimeters" that probably means it is closer to 200.00 mm than to 200.05 mm or 199.95 mm.

We all know that the acceleration due to gravity varies from place to place on the earth's surface. It also varies with the height above the surface, and gravity meters capable of measuring the variation from the floor to a tabletop are readily available. Further, any physical measure such as can only be determined by means of an experiment, and since a perfect experimental apparatus does not exist, it is impossible even in principle to ever know perfectly. Thus, the specification of given above is useful only as a possible exercise for a student. In order to give it some meaning it must be changed to something like:

Two questions arise about the measurement. First, is it "accurate," in other words, did the experiment work properly and were all the necessary factors taken into account? The answer to this depends on the skill of the experimenter in identifying and eliminating all systematic errors. These are discussed in Section 3.4.

The second question regards the "precision" of the experiment. In this case the precision of the result is given: the experimenter claims the precision of the result is within 0.03 m/s

1. The person who did the measurement probably had some "gut feeling" for the precision and "hung" an error on the result primarily to communicate this feeling to other people. Common sense should always take precedence over mathematical manipulations.

2. In complicated experiments, error analysis can identify dominant errors and hence provide a guide as to where more effort is needed to improve an experiment.

3. There is virtually no case in the experimental physical sciences where the correct error analysis is to compare the result with a number in some book. A correct experiment is one that is performed correctly, not one that gives a result in agreement with other measurements.

4. The best precision possible for a given experiment is always limited by the apparatus. Polarization measurements in high-energy physics require tens of thousands of person-hours and cost hundreds of thousand of dollars to perform, and a good measurement is within a factor of two. Electrodynamics experiments are considerably cheaper, and often give results to 8 or more significant figures. In both cases, the experimenter must struggle with the equipment to get the most precise and accurate measurement possible.

3.1.2 Different Types of Errors

As mentioned above, there are two types of errors associated with an experimental result: the "precision" and the "accuracy". One well-known text explains the difference this way:

" " E.M. Pugh and G.H. Winslow, p. 6.

The object of a good experiment is to minimize both the errors of precision and the errors of accuracy.

Usually, a given experiment has one or the other type of error dominant, and the experimenter devotes the most effort toward reducing that one. For example, in measuring the height of a sample of geraniums to determine an average value, the random variations within the sample of plants are probably going to be much larger than any possible inaccuracy in the ruler being used. Similarly for many experiments in the biological and life sciences, the experimenter worries most about increasing the precision of his/her measurements. Of course, some experiments in the biological and life sciences are dominated by errors of accuracy.

On the other hand, in titrating a sample of HCl acid with NaOH base using a phenolphthalein indicator, the major error in the determination of the original concentration of the acid is likely to be one of the following: (1) the accuracy of the markings on the side of the burette; (2) the transition range of the phenolphthalein indicator; or (3) the skill of the experimenter in splitting the last drop of NaOH. Thus, the accuracy of the determination is likely to be much worse than the precision. This is often the case for experiments in chemistry, but certainly not all.

Question: Most experiments use theoretical formulas, and usually those formulas are approximations. Is the error of approximation one of precision or of accuracy?

3.1.3 References

There is extensive literature on the topics in this chapter. The following lists some well-known introductions.

D.C. Baird, (Prentice-Hall, 1962)

E.M. Pugh and G.H. Winslow, (Addison-Wesley, 1966)

J.R. Taylor, (University Science Books, 1982)

In addition, there is a web document written by the author of that is used to teach this topic to first year Physics undergraduates at the University of Toronto. The following Hyperlink points to that document.

3.2 Determining the Precision

3.2.1 The Standard Deviation

In the nineteenth century, Gauss' assistants were doing astronomical measurements. However, they were never able to exactly repeat their results. Finally, Gauss got angry and stormed into the lab, claiming he would show these people how to do the measurements once and for all. The only problem was that Gauss wasn't able to repeat his measurements exactly either!

After he recovered his composure, Gauss made a histogram of the results of a particular measurement and discovered the famous Gaussian or bell-shaped curve.

Many people's first introduction to this shape is the grade distribution for a course. Here is a sample of such a distribution, using the function .

We use a standard package to generate a Probability Distribution Function ( ) of such a "Gaussian" or "normal" distribution. The mean is chosen to be 78 and the standard deviation is chosen to be 10; both the mean and standard deviation are defined below.

We then normalize the distribution so the maximum value is close to the maximum number in the histogram and plot the result.

In this graph,

Finally, we look at the histogram and plot together.

We can see the functional form of the Gaussian distribution by giving symbolic values.

In this formula, the quantity , and . The is sometimes called the . The definition of is as follows.

Here is the total number of measurements and is the result of measurement number .

The standard deviation is a measure of the width of the peak, meaning that a larger value gives a wider peak.

If we look at the area under the curve from graph, we find that this area is 68 percent of the total area. Thus, any result chosen at random has a 68% change of being within one standard deviation of the mean. We can show this by evaluating the integral. For convenience, we choose the mean to be zero.

Now, we numericalize this and multiply by 100 to find the percent.

The only problem with the above is that the measurement must be repeated an infinite number of times before the standard deviation can be determined. If is less than infinity, one can only estimate measurements, this is the best estimate.

The major difference between this estimate and the definition is the . This is reasonable since if = 1 we know we can't determine

Here is an example. Suppose we are to determine the diameter of a small cylinder using a micrometer. We repeat the measurement 10 times along various points on the cylinder and get the following results, in centimeters.

The number of measurements is the length of the list.

The average or mean is now calculated.

Then the standard deviation is to be 0.00185173.

We repeat the calculation in a functional style.

Note that the package, which is standard with , includes functions to calculate all of these quantities and a great deal more.

We close with two points:

1. The standard deviation has been associated with the error in each individual measurement. Section 3.3.2 discusses how to find the error in the estimate of the average.

2. This calculation of the standard deviation is only an estimate. In fact, we can find the expected error in the estimate,

As discussed in more detail in Section 3.3, this means that the true standard deviation probably lies in the range of values.

Viewed in this way, it is clear that the last few digits in the numbers above for function adjusts these significant figures based on the error.

is discussed further in Section 3.3.1.

3.2.2 The Reading Error

There is another type of error associated with a directly measured quantity, called the "reading error". Referring again to the example of Section 3.2.1, the measurements of the diameter were performed with a micrometer. The particular micrometer used had scale divisions every 0.001 cm. However, it was possible to estimate the reading of the micrometer between the divisions, and this was done in this example. But, there is a reading error associated with this estimation. For example, the first data point is 1.6515 cm. Could it have been 1.6516 cm instead? How about 1.6519 cm? There is no fixed rule to answer the question: the person doing the measurement must guess how well he or she can read the instrument. A reasonable guess of the reading error of this micrometer might be 0.0002 cm on a good day. If the experimenter were up late the night before, the reading error might be 0.0005 cm.

An important and sometimes difficult question is whether the reading error of an instrument is "distributed randomly". Random reading errors are caused by the finite precision of the experiment. If an experimenter consistently reads the micrometer 1 cm lower than the actual value, then the reading error is not random.

For a digital instrument, the reading error is ± one-half of the last digit. Note that this assumes that the instrument has been properly engineered to round a reading correctly on the display.

3.2.3 "THE" Error

So far, we have found two different errors associated with a directly measured quantity: the standard deviation and the reading error. So, which one is the actual real error of precision in the quantity? The answer is both! However, fortunately it almost always turns out that one will be larger than the other, so the smaller of the two can be ignored.

In the diameter example being used in this section, the estimate of the standard deviation was found to be 0.00185 cm, while the reading error was only 0.0002 cm. Thus, we can use the standard deviation estimate to characterize the error in each measurement. Another way of saying the same thing is that the observed spread of values in this example is not accounted for by the reading error. If the observed spread were more or less accounted for by the reading error, it would not be necessary to estimate the standard deviation, since the reading error would be the error in each measurement.

Of course, everything in this section is related to the precision of the experiment. Discussion of the accuracy of the experiment is in Section 3.4.

3.2.4 Rejection of Measurements

Often when repeating measurements one value appears to be spurious and we would like to throw it out. Also, when taking a series of measurements, sometimes one value appears "out of line". Here we discuss some guidelines on rejection of measurements; further information appears in Chapter 7.

It is important to emphasize that the whole topic of rejection of measurements is awkward. Some scientists feel that the rejection of data is justified unless there is evidence that the data in question is incorrect. Other scientists attempt to deal with this topic by using quasi-objective rules such as 's . Still others, often incorrectly, throw out any data that appear to be incorrect. In this section, some principles and guidelines are presented; further information may be found in many references.

First, we note that it is incorrect to expect each and every measurement to overlap within errors. For example, if the error in a particular quantity is characterized by the standard deviation, we only expect 68% of the measurements from a normally distributed population to be within one standard deviation of the mean. Ninety-five percent of the measurements will be within two standard deviations, 99% within three standard deviations, etc., but we never expect 100% of the measurements to overlap within any finite-sized error for a truly Gaussian distribution.

Of course, for most experiments the assumption of a Gaussian distribution is only an approximation.

If the error in each measurement is taken to be the reading error, again we only expect most, not all, of the measurements to overlap within errors. In this case the meaning of "most", however, is vague and depends on the optimism/conservatism of the experimenter who assigned the error.

Thus, it is always dangerous to throw out a measurement. Maybe we are unlucky enough to make a valid measurement that lies ten standard deviations from the population mean. A valid measurement from the tails of the underlying distribution should not be thrown out. It is even more dangerous to throw out a suspect point indicative of an underlying physical process. Very little science would be known today if the experimenter always threw out measurements that didn't match preconceived expectations!

In general, there are two different types of experimental data taken in a laboratory and the question of rejecting measurements is handled in slightly different ways for each. The two types of data are the following:

1. A series of measurements taken with one or more variables changed for each data point. An example is the calibration of a thermocouple, in which the output voltage is measured when the thermocouple is at a number of different temperatures.

2. Repeated measurements of the same physical quantity, with all variables held as constant as experimentally possible. An example is the measurement of the height of a sample of geraniums grown under identical conditions from the same batch of seed stock.

For a series of measurements (case 1), when one of the data points is out of line the natural tendency is to throw it out. But, as already mentioned, this means you are assuming the result you are attempting to measure. As a rule of thumb, unless there is a physical explanation of why the suspect value is spurious and it is no more than three standard deviations away from the expected value, it should probably be kept. Chapter 7 deals further with this case.

For repeated measurements (case 2), the situation is a little different. Say you are measuring the time for a pendulum to undergo 20 oscillations and you repeat the measurement five times. Assume that four of these trials are within 0.1 seconds of each other, but the fifth trial differs from these by 1.4 seconds ( , more than three standard deviations away from the mean of the "good" values). There is no known reason why that one measurement differs from all the others. Nonetheless, you may be justified in throwing it out. Say that, unknown to you, just as that measurement was being taken, a gravity wave swept through your region of spacetime. However, if you are trying to measure the period of the pendulum when there are no gravity waves affecting the measurement, then throwing out that one result is reasonable. (Although trying to repeat the measurement to find the existence of gravity waves will certainly be more fun!) So whatever the reason for a suspect value, the rule of thumb is that it may be thrown out provided that fact is well documented and that the measurement is repeated a number of times more to convince the experimenter that he/she is not throwing out an important piece of data indicating a new physical process.

3.3 Propagation of Errors of Precision

3.3.1 Discussion and Examples

Usually, errors of precision are probabilistic. This means that the experimenter is saying that the actual value of some parameter is within a specified range. For example, if the half-width of the range equals one standard deviation, then the probability is about 68% that over repeated experimentation the true mean will fall within the range; if the half-width of the range is twice the standard deviation, the probability is 95%, etc.

If we have two variables, say and , and want to combine them to form a new variable, we want the error in the combination to preserve this probability.

The correct procedure to do this is to combine errors in quadrature, which is the square root of the sum of the squares. supplies a function.

For simple combinations of data with random errors, the correct procedure can be summarized in three rules. will stand for the errors of precision in , , and , respectively. We assume that and are independent of each other.

Note that all three rules assume that the error, say , is small compared to the value of .

If

z = x * y

or

then

In words, the fractional error in is the quadrature of the fractional errors in and .

If

z = x + y

or

z = x - y

then

In words, the error in is the quadrature of the errors in and .

If

then

or equivalently

includes functions to combine data using the above rules. They are named , , , , and .

Imagine we have pressure data, measured in centimeters of Hg, and volume data measured in arbitrary units. Each data point consists of { , } pairs.

We calculate the pressure times the volume.

In the above, the values of and have been multiplied and the errors have ben combined using Rule 1.

There is an equivalent form for this calculation.

Consider the first of the volume data: {11.28156820762763, 0.031}. The error means that the true value is claimed by the experimenter to probably lie between 11.25 and 11.31. Thus, all the significant figures presented to the right of 11.28 for that data point really aren't significant. The function will adjust the volume data.

Notice that by default, uses the two most significant digits in the error for adjusting the values. This can be controlled with the option.

For most cases, the default of two digits is reasonable. As discussed in Section 3.2.1, if we assume a normal distribution for the data, then the fractional error in the determination of the standard deviation , and can be written as follows.

Thus, using this as a general rule of thumb for all errors of precision, the estimate of the error is only good to 10%, ( one significant figure, unless is greater than 51) . Nonetheless, keeping two significant figures handles cases such as 0.035 vs. 0.030, where some significance may be attached to the final digit.

You should be aware that when a datum is massaged by , the extra digits are dropped.

By default, and the other functions use the function. The use of is controlled using the option.

The number of digits can be adjusted.

To form a power, say,

we might be tempted to just do

function.

Finally, imagine that for some reason we wish to form a combination.

We might be tempted to solve this with the following.

then the error is

Here is an example solving . We shall use and below to avoid overwriting the symbols and . First we calculate the total derivative.

Next we form the error.

Now we can evaluate using the pressure and volume data to get a list of errors.

Next we form the list of pairs.

The function combines these steps with default significant figure adjustment.

The function can be used in place of the other functions discussed above.

In this example, the function will be somewhat faster.

There is a caveat in using . The expression must contain only symbols, numerical constants, and arithmetic operations. Otherwise, the function will be unable to take the derivatives of the expression necessary to calculate the form of the error. The other functions have no such limitation.

3.3.1.1 Another Approach to Error Propagation: The and Datum

value error

Data[{{789.7, 2.2}, {790.8, 2.3}, {791.2, 2.3}, {792.6, 2.4}, {791.8, 2.5},
{792.2, 2.5}, {794.7, 2.6}, {794., 2.6}, {794.4, 2.7}, {795.3, 2.8},
{796.4, 2.8}}]Data[{{789.7, 2.2}, {790.8, 2.3}, {791.2, 2.3}, {792.6, 2.4}, {791.8, 2.5},

{792.2, 2.5}, {794.7, 2.6}, {794., 2.6}, {794.4, 2.7}, {795.3, 2.8},

{796.4, 2.8}}]

The wrapper can be removed.

{{789.7, 2.2}, {790.8, 2.3}, {791.2, 2.3}, {792.6, 2.4}, {791.8, 2.5},
{792.2, 2.5}, {794.7, 2.6}, {794., 2.6}, {794.4, 2.7}, {795.3, 2.8}, {796.4, 2.8}}{{789.7, 2.2}, {790.8, 2.3}, {791.2, 2.3}, {792.6, 2.4}, {791.8, 2.5},

{792.2, 2.5}, {794.7, 2.6}, {794., 2.6}, {794.4, 2.7}, {795.3, 2.8}, {796.4, 2.8}}

The reason why the output of the previous two commands has been formatted as is that typesets the pairs using ± for output.

A similar construct can be used with individual data points.

Datum[{70, 0.04}]Datum[{70, 0.04}]

Just as for , the typesetting of uses

The and constructs provide "automatic" error propagation for multiplication, division, addition, subtraction, and raising to a power. Another advantage of these constructs is that the rules built into know how to combine data with constants.

The rules also know how to propagate errors for many transcendental functions.

This rule assumes that the error is small relative to the value, so we can approximate.

or arguments, are given by .

We have seen that typesets the and constructs using ±. The function can be used directly, and provided its arguments are numeric, errors will be propagated.

One may typeset the ± into the input expression, and errors will again be propagated.

The ± input mechanism can combine terms by addition, subtraction, multiplication, division, raising to a power, addition and multiplication by a constant number, and use of the . The rules used by for ± are only for numeric arguments.

This makes different than

3.3.1.2 Why Quadrature?

Here we justify combining errors in quadrature. Although they are not proofs in the usual pristine mathematical sense, they are correct and can be made rigorous if desired.

First, you may already know about the "Random Walk" problem in which a player starts at the point = 0 and at each move steps either forward (toward + ) or backward (toward - ). The choice of direction is made randomly for each move by, say, flipping a coin. If each step covers a distance , then after steps the expected most probable distance of the player from the origin can be shown to be

Thus, the distance goes up as the square root of the number of steps.

Now consider a situation where measurements of a quantity are performed, each with an identical random error . We find the sum of the measurements.

, it is equally likely to be + as - , and which is essentially random. Thus, the expected most probable error in the sum goes up as the square root of the number of measurements.

This is exactly the result obtained by combining the errors in quadrature.

Another similar way of thinking about the errors is that in an abstract linear error space, the errors span the space. If the errors are probabilistic and uncorrelated, the errors in fact are linearly independent (orthogonal) and thus form a basis for the space. Thus, we would expect that to add these independent random errors, we would have to use Pythagoras' theorem, which is just combining them in quadrature.

3.3.2 Finding the Error in an Average

The rules for propagation of errors, discussed in Section 3.3.1, allow one to find the error in an average or mean of a number of repeated measurements. Recall that to compute the average, first the sum of all the measurements is found, and the rule for addition of quantities allows the computation of the error in the sum. Next, the sum is divided by the number of measurements, and the rule for division of quantities allows the calculation of the error in the result ( the error of the mean).

In the case that the error in each measurement has the same value, the result of applying these rules for propagation of errors can be summarized as a theorem.

Theorem: If the measurement of a random variable is repeated times, and the random variable has standard deviation , then the standard deviation in the mean is

Proof: One makes measurements, each with error .

{x1, errx}, {x2, errx}, ... , {xn, errx}

We calculate the sum.

sumx = x1 + x2 + ... + xn

We calculate the error in the sum.

This last line is the key: by repeating the measurements times, the error in the sum only goes up as [ ].

The mean

Applying the rule for division we get the following.

This completes the proof.

The quantity called

Here is an example. In Section 3.2.1, 10 measurements of the diameter of a small cylinder were discussed. The mean of the measurements was 1.6514 cm and the standard deviation was 0.00185 cm. Now we can calculate the mean and its error, adjusted for significant figures.

Note that presenting this result without significant figure adjustment makes no sense.

The above number implies that there is meaning in the one-hundred-millionth part of a centimeter.

Here is another example. Imagine you are weighing an object on a "dial balance" in which you turn a dial until the pointer balances, and then read the mass from the marking on the dial. You find = 26.10 ± 0.01 g. The 0.01 g is the reading error of the balance, and is about as good as you can read that particular piece of equipment. You remove the mass from the balance, put it back on, weigh it again, and get = 26.10 ± 0.01 g. You get a friend to try it and she gets the same result. You get another friend to weigh the mass and he also gets = 26.10 ± 0.01 g. So you have four measurements of the mass of the body, each with an identical result. Do you think the theorem applies in this case? If yes, you would quote = 26.100 ± 0.01/ [4] = 26.100 ± 0.005 g. How about if you went out on the street and started bringing strangers in to repeat the measurement, each and every one of whom got = 26.10 ± 0.01 g. So after a few weeks, you have 10,000 identical measurements. Would the error in the mass, as measured on that $50 balance, really be the following?

The point is that these rules of statistics are only a rough guide and in a situation like this example where they probably don't apply, don't be afraid to ignore them and use your "uncommon sense". In this example, presenting your result as = 26.10 ± 0.01 g is probably the reasonable thing to do.

3.4 Calibration, Accuracy, and Systematic Errors

In Section 3.1.2, we made the distinction between errors of precision and accuracy by imagining that we had performed a timing measurement with a very precise pendulum clock, but had set its length wrong, leading to an inaccurate result. Here we discuss these types of errors of accuracy. To get some insight into how such a wrong length can arise, you may wish to try comparing the scales of two rulers made by different companies — discrepancies of 3 mm across 30 cm are common!

If we have access to a ruler we trust ( a "calibration standard"), we can use it to calibrate another ruler. One reasonable way to use the calibration is that if our instrument measures and the standard records , then we can multiply all readings of our instrument by / . Since the correction is usually very small, it will practically never affect the error of precision, which is also small. Calibration standards are, almost by definition, too delicate and/or expensive to use for direct measurement.

Here is an example. We are measuring a voltage using an analog Philips multimeter, model PM2400/02. The result is 6.50 V, measured on the 10 V scale, and the reading error is decided on as 0.03 V, which is 0.5%. Repeating the measurement gives identical results. It is calculated by the experimenter that the effect of the voltmeter on the circuit being measured is less than 0.003% and hence negligible. However, the manufacturer of the instrument only claims an accuracy of 3% of full scale (10 V), which here corresponds to 0.3 V.

Now, what this claimed accuracy means is that the manufacturer of the instrument claims to control the tolerances of the components inside the box to the point where the value read on the meter will be within 3% times the scale of the actual value. Furthermore, this is not a random error; a given meter will supposedly always read too high or too low when measurements are repeated on the same scale. Thus, repeating measurements will not reduce this error.

A further problem with this accuracy is that while most good manufacturers (including Philips) tend to be quite conservative and give trustworthy specifications, there are some manufacturers who have the specifications written by the sales department instead of the engineering department. And even Philips cannot take into account that maybe the last person to use the meter dropped it.

Nonetheless, in this case it is probably reasonable to accept the manufacturer's claimed accuracy and take the measured voltage to be 6.5 ± 0.3 V. If you want or need to know the voltage better than that, there are two alternatives: use a better, more expensive voltmeter to take the measurement or calibrate the existing meter.

Using a better voltmeter, of course, gives a better result. Say you used a Fluke 8000A digital multimeter and measured the voltage to be 6.63 V. However, you're still in the same position of having to accept the manufacturer's claimed accuracy, in this case (0.1% of reading + 1 digit) = 0.02 V. To do better than this, you must use an even better voltmeter, which again requires accepting the accuracy of this even better instrument and so on, ad infinitum, until you run out of time, patience, or money.

Say we decide instead to calibrate the Philips meter using the Fluke meter as the calibration standard. Such a procedure is usually justified only if a large number of measurements were performed with the Philips meter. Why spend half an hour calibrating the Philips meter for just one measurement when you could use the Fluke meter directly?

We measure four voltages using both the Philips and the Fluke meter. For the Philips instrument we are not interested in its accuracy, which is why we are calibrating the instrument. So we will use the reading error of the Philips instrument as the error in its measurements and the accuracy of the Fluke instrument as the error in its measurements.

We form lists of the results of the measurements.

We can examine the differences between the readings either by dividing the Fluke results by the Philips or by subtracting the two values.

The second set of numbers is closer to the same value than the first set, so in this case adding a correction to the Philips measurement is perhaps more appropriate than multiplying by a correction.

We form a new data set of format { }.

We can guess, then, that for a Philips measurement of 6.50 V the appropriate correction factor is 0.11 ± 0.04 V, where the estimated error is a guess based partly on a fear that the meter's inaccuracy may not be as smooth as the four data points indicate. Thus, the corrected Philips reading can be calculated.

(You may wish to know that all the numbers in this example are real data and that when the Philips meter read 6.50 V, the Fluke meter measured the voltage to be 6.63 ± 0.02 V.)

Finally, a further subtlety: Ohm's law states that the resistance is related to the voltage and the current across the resistor according to the following equation.

V = IR

Imagine that we are trying to determine an unknown resistance using this law and are using the Philips meter to measure the voltage. Essentially the resistance is the slope of a graph of voltage versus current.

If the Philips meter is systematically measuring all voltages too big by, say, 2%, that systematic error of accuracy will have no effect on the slope and therefore will have no effect on the determination of the resistance . So in this case and for this measurement, we may be quite justified in ignoring the inaccuracy of the voltmeter entirely and using the reading error to determine the uncertainty in the determination of .

3.5 Summary of the Error Propagation Routines

  • Wolfram|Alpha Notebook Edition
  • Mobile Apps
  • Wolfram Workbench
  • Volume & Site Licensing
  • View all...
  • For Customers
  • Online Store
  • Product Registration
  • Product Downloads
  • Service Plans Benefits
  • User Portal
  • Your Account
  • Customer Service
  • Get Started with Wolfram
  • Fast Introduction for Math Students
  • Public Resources
  • Wolfram|Alpha
  • Resource System
  • Connected Devices Project
  • Wolfram Data Drop
  • Wolfram Science
  • Computational Thinking
  • About Wolfram
  • Legal & Privacy Policy

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 11 November 2022

Error, reproducibility and uncertainty in experiments for electrochemical energy technologies

  • Graham Smith   ORCID: orcid.org/0000-0003-0713-2893 1 &
  • Edmund J. F. Dickinson   ORCID: orcid.org/0000-0003-2137-3327 1  

Nature Communications volume  13 , Article number:  6832 ( 2022 ) Cite this article

10k Accesses

9 Citations

22 Altmetric

Metrics details

  • Electrocatalysis
  • Electrochemistry
  • Materials for energy and catalysis

The authors provide a metrology-led perspective on best practice for the electrochemical characterisation of materials for electrochemical energy technologies. Such electrochemical experiments are highly sensitive, and their results are, in practice, often of uncertain quality and challenging to reproduce quantitatively.

A critical aspect of research on electrochemical energy devices, such as batteries, fuel cells and electrolysers, is the evaluation of new materials, components, or processes in electrochemical cells, either ex situ, in situ or in operation. For such experiments, rigorous experimental control and standardised methods are required to achieve reproducibility, even on standard or idealised systems such as single crystal platinum 1 . Data reported for novel materials often exhibit high (or unstated) uncertainty and often prove challenging to reproduce quantitatively. This situation is exacerbated by a lack of formally standardised methods, and practitioners with less formal training in electrochemistry being unaware of best practices. This limits trust in published metrics, with discussions on novel electrochemical systems frequently focusing on a single series of experiments performed by one researcher in one laboratory, comparing the relative performance of the novel material against a claimed state-of-the-art.

Much has been written about the broader reproducibility/replication crisis 2 and those reading the electrochemical literature will be familiar with weakly underpinned claims of “outstanding” performance, while being aware that comparisons may be invalidated by measurement errors introduced by experimental procedures which violate best practice; such issues frequently mar otherwise exciting science in this area. The degree of concern over the quality of reported results is evidenced by the recent decision of several journals to publish explicit experimental best practices 3 , 4 , 5 , reporting guidelines or checklists 6 , 7 , 8 , 9 , 10 and commentary 11 , 12 , 13 aiming to improve the situation, including for parallel theoretical work 14 .

We write as two electrochemists who, working in a national metrology institute, have enjoyed recent exposure to metrology: the science of measurement. Metrology provides the vocabulary 15 and mathematical tools 16 to express confidence in measurements and the comparisons made between them. Metrological systems and frameworks for quantification underpin consistency and assurance in all measurement fields and formal metrology is an everyday consideration for practical and academic work in fields where accurate measurements are crucial; we have found it a useful framework within which to evaluate our own electrochemical work. Here, rather than pen another best practice guide, we aim, with focus on three-electrode electrochemical measurements for energy material characterisation, to summarise some advice that we hope helps those performing electrochemical experiments to:

avoid mistakes and minimise error

report in a manner that facilitates reproducibility

consider and quantify uncertainty

Minimising mistakes and error

Metrology dispenses with nebulous concepts such as performance and instead requires scientists to define a specific measurand (“the quantity intended to be measured”) along with a measurement model ( ”the mathematical relation among all quantities known to be involved in a measurement”), which converts the experimental indicators into the measurand 15 . Error is the difference between the reported value of this measurand and its unknowable true value. (Note this is not the formal definition, and the formal concepts of error and true value are not fully compatible with measurement concepts discussed in this article, but we retain it here—as is common in metrology tuition delivered by national metrology institutes—for pedagogical purposes 15 ).

Mistakes (or gross errors) are those things which prevent measurements from working as intended. In electrochemistry the primary experimental indicator is often current or voltage, while the measurand might be something simple, like device voltage for a given current density, or more complex, like a catalyst’s turnover frequency. Both of these are examples of ‘method-defined measurands’, where the results need to be defined in reference to the method of measurement 17 , 18 (for example, to take account of operating conditions). Robust experimental design and execution are vital to understand, quantify and minimise sources of error, and to prevent mistakes.

Contemporary electrochemical instrumentation can routinely offer a current resolution and accuracy on the order of femtoamps; however, one electron looks much like another to a potentiostat. Consequently, the practical limit on measurements of current is the scientist’s ability to unambiguously determine what causes the observed current. Crucially, they must exclude interfering processes such as modified/poisoned catalyst sites or competing reactions due to impurities.

As electrolytes are conventionally in enormous excess compared to the active heterogeneous interface, electrolyte purity requirements are very high. Note, for example, that a perfectly smooth 1 cm 2 polycrystalline platinum electrode has on the order of 2 nmol of atoms exposed to the electrolyte, so that irreversibly adsorbing impurities present at the part per billion level (nmol mol −1 ) in the electrolyte may substantially alter the surface of the electrode. Sources of impurities at such low concentration are innumerable and must be carefully considered for each experiment; impurity origins for kinetic studies in aqueous solution have been considered broadly in the historical literature, alongside a review of standard mitigation methods 19 . Most commercial electrolytes contain impurities and the specific ‘grade’ chosen may have a large effect; for example, one study showed a three-fold decrease in the specific activity of oxygen reduction catalysts when preparing electrolytes with American Chemical Society (ACS) grade acid rather than a higher purity grade 20 . Likewise, even 99.999% pure hydrogen gas, frequently used for sparging, may contain more than the 0.2 μmol mol −1 of carbon monoxide permitted for fuel cell use 21 .

The most insidious impurities are those generated in situ. The use of reference electrodes with chloride-containing filling solutions should be avoided where chloride may poison catalysts 22 or accelerate dissolution. Similarly, reactions at the counter electrode, including dissolution of the electrode itself, may result in impurities. This is sometimes overlooked when platinum counter electrodes are used to assess ‘platinum-free’ electrocatalysts, accidentally resulting in performance-enhancing contamination 23 , 24 ; a critical discussion on this topic has recently been published 25 . Other trace impurity sources include plasticisers present in cells and gaskets, or silicates from the inappropriate use of glass when working with alkaline electrolytes 26 . To mitigate sensitivity to impurities from the environment, cleaning protocols for cells and components must be robust 27 . The use of piranha solution or similarly oxidising solution followed by boiling in Type 1 water is typical when performing aqueous electrochemistry 20 . Cleaned glassware and electrodes are also routinely stored underwater to prevent recontamination from airborne impurities.

The behaviour of electronic hardware used for electrochemical experiments should be understood and considered carefully in interpreting data 28 , recognising that the built-in complexity of commercially available digital potentiostats (otherwise advantageous!) is capable of introducing measurement artefacts or ambiguity 29 , 30 . While contemporary electrochemical instrumentation may have a voltage resolution of ~1 μV, its voltage measurement uncertainty is limited by other factors, and is typically on the order of 1 mV. As passing current through an electrode changes its potential, a dedicated reference electrode is often incorporated into both ex situ and, increasingly, in situ experiments to provide a stable well defined reference. Reference electrodes are typically selected from a range of well-known standardised electrode–electrolyte interfaces at which a characteristic and kinetically rapid reversible faradaic process occurs. The choice of reference electrode should be made carefully in consideration of chemical compatibility with the measurement environment 31 , 32 , 33 , 34 . In combination with an electronic blocking resistance, the potential of the electrode should be stable and reproducible. Unfortunately, deviation from the ideal behaviour frequently occurs. While this can often be overlooked when comparing results from identical cells, more attention is required when reporting values for comparison.

In all cases where conversion between different electrolyte–reference electrode systems is required, junction potentials should be considered. These arise whenever there are different chemical conditions in the electrolyte at the working electrode and reference electrode interfaces. Outside highly dilute solutions, or where there are large activity differences for a reactant/product of the electrode reaction (e.g. pH for hydrogen reactions), liquid junction potentials for conventional aqueous ions have been estimated in the range <50 mV 33 . Such a deviation may nonetheless be significant when onset potentials or activities at specific potentials are being reported. The measured potential difference between the working and reference electrode also depends strongly on the geometry of the cell, so cell design is critical. Fig.  1 shows the influence of cell design on potential profiles. Ideally the reference electrode should therefore be placed close to the working electrode (noting that macroscopic electrodes may have inhomogeneous potentials). To minimise shielding of the electric field between counter and working electrode and interruption of mass transport processes, a thin Luggin-Haber capillary is often used and a small separation maintained. Understanding of shielding and edge effects is vital when reference electrodes are introduced in situ. This is especially applicable for analysis of energy devices for which constraints on cell design, due to the need to minimise electrolyte resistance and seal the cell, preclude optimal reference electrode positioning 32 , 35 , 36 .

figure 1

a Illustration (simulated data) of primary (resistive) current and potential distribution in a typical three-electrode cell. The main compartment is cylindrical (4 cm diameter, 1 cm height), filled with electrolyte with conductivity 1.28 S m −1 (0.1 M KCl(aq)). The working electrode (WE) is a 2 mm diameter disc drawing 1 mA (≈ 32 mA cm −2 ) from a faradaic process with infinitely fast kinetics and redox potential 0.2 V vs the reference electrode (RE). The counter electrode (CE) is connected to the main compartment by a porous frit; the RE is connected by a Luggin capillary (green cylinders) whose tip position is offset from the WE by a variable distance. Red lines indicate prevailing current paths; coloured surfaces indicate isopotential contours normal to the current density. b Plot of indicated WE vs RE potential (simulated data). As the Luggin tip is moved away from the WE surface, ohmic losses due to the WE-CE current distribution lead to variation in the indicated WE-RE potential. Appreciable error may arise on an offset length scale comparable to the WE radius.

Quantitative statements about fundamental electrochemical processes based on measured values of current and voltage inevitably rely on models of the system. Such models have assumptions that may be routinely overlooked when following experimental and analysis methods, and that may restrict their application to real-world systems. It is quite possible to make highly precise but meaningless measurements! An often-assumed condition for electrocatalyst analysis is the absence of mass transport limitation. For some reactions, such as the acidic hydrogen oxidation and hydrogen evolution reactions, this state is arguably so challenging to reach at representative conditions that it is impossible to measure true catalyst activity 11 . For example, ex situ thin-film rotating disk electrode measurements routinely fail to predict correct trends in catalyst performance in morphologically complex catalyst layers  as relevant operating conditions (e.g. meaningful current densities) are theoretically inaccessible. This topic has been extensively discussed with some authors directly criticising this technique and exploring alternatives 37 , 38 , and others defending the technique’s applicability for ranking catalysts if scrupulous attention is paid to experimental details 39 ; yet, many reports continue to use this measurement technique blindly with no regard for its applicability. We therefore strongly urge those planning measurements to consider whether their chosen technique is capable of providing sufficient evidence to disprove their hypothesis, even if it has been widely used for similar experiments.

The correct choice of technique should be dependent upon the measurand being probed rather than simply following previous reports. The case of iR correction, where a measurement of the uncompensated resistance is used to correct the applied voltage, is a good example. When the measurand is a material property, such as intrinsic catalyst activity, the uncompensated resistance is a source of error introduced by the experimental method and it should carefully be corrected out (Fig.  1 ). In the case that the uncompensated resistance is intrinsic to the measurand—for instance the operating voltage of an electrolyser cell—iR compensation is inappropriate and only serves to obfuscate. Another example is the choice of ex situ (outside the operating environment), in situ (in the operating environment), and operando (during operation) measurements. While in situ or operando testing allows characterisation under conditions that are more representative of real-world use, it may also yield measurements with increased uncertainty due to the decreased possibility for fine experimental control. Depending on the intended use of the measurement, an informed compromise must be sought between how relevant and how uncertain the resulting measurement will be.

Maximising reproducibility

Most electrochemists assess the repeatability of measurements, performing the same measurement themselves several times. Repeats, where all steps (including sample preparation, where relevant) of a measurement are carried out multiple times, are absolutely crucial for highlighting one-off mistakes (Fig.  2 ). Reproducibility, however, is assessed when comparing results reported by different laboratories. Many readers will be familiar with the variability in key properties reported for various systems e.g. variability in the reported electrochemically active surface area (ECSA) of commercial catalysts, which might reasonably be expected to be constant, suggesting that, in practice, the reproducibility of results cannot be taken for granted. As electrochemistry deals mostly with method-defined measurands, the measurement procedure must be standardised for results to be comparable. Variation in results therefore strongly suggests that measurements are not being performed consistently and that the information typically supplied when publishing experimental methods is insufficient to facilitate reproducibility of electrochemical measurements. Quantitative electrochemical measurements require control over a large range of parameters, many of which are easily overlooked or specified imprecisely when reporting data. An understanding of the crucial parameters and methods for their control is often institutional knowledge, held by expert electrochemists, but infrequently formalised and communicated e.g. through publication of widely adopted standards. This creates challenges to both reproducibility and the corresponding assessment of experimental quality by reviewers. The reporting standards established by various publishers (see Introduction) offer a practical response, but it is still unclear whether these will contain sufficiently granular detail to improve the situation.

figure 2

The measurements from laboratory 1 show a high degree of repeatability, while the measurements from laboratory 2 do not. Apparently, a mistake has been made in repeat 1, which will need to be excluded from any analysis and any uncertainty analysis, and/or suggests further repeat measurements should be conducted. The error bars are based on an uncertainty with coverage factor ~95% (see below) so the results from the two laboratories are different, i.e. show poor reproducibility. This may indicate differing experimental practice or that some as yet unidentified parameter is influencing the results.

Besides information typically supplied in the description of experimental methods for publication, which, at a minimum, must detail the materials, equipment and measurement methods used to generate the results, we suggest that a much more comprehensive description is often required, especially where measurements have historically poor reproducibility or the presented results differ from earlier reports. Such an expanded ‘supplementary experimental’ section would additionally include any details that could impact the results: for example, material pre-treatment, detailed electrode preparation steps, cleaning procedures, expected electrolyte and gas impurities, electrode preconditioning processes, cell geometry including electrode positions, detail of junctions between electrodes, and any other fine experimental details which might be institutional knowledge but unknown to the (now wide) readership of the electrochemical literature. In all cases any corrections and calculations used should be specified precisely and clearly justified; these may include determinations of properties of the studied system, such as ECSA, or of the environment, such as air pressure. We highlight that knowledge of the ECSA is crucial for conventional reporting of intrinsic electrocatalyst activity, but is often very challenging to measure in a reproducible manner 40 , 41 .

To aid reproducibility we recommend regularly calibrating experimental equipment and doing so in a way that is traceable to primary standards realising the International System of Units (SI) base units. The SI system ensures that measurement units (such as the volt) are uniform globally and invariant over time. Calibration applies to direct experimental indicators, e.g. loads and potentiostats, but equally to supporting tools such as temperature probes, balances, and flow meters. Calibration of reference electrodes is often overlooked even though variations from ideal behaviour can be significant 42 and, as discussed above, are often the limit of accuracy on potential measurement. Sometimes reports will specify internal calibration against a known reaction (such as the onset of the hydrogen evolution reaction), but rarely detail regular comparisons to a local master electrode artefact such as a reference hydrogen electrode or explain how that artefact is traceable, e.g. through control of the filling solution concentration and measurement conditions. If reference is made to a standardised material (e.g. commercial Pt/C) the specified material should be both widely available and the results obtained should be consistent with prior reports.

Beyond calibration and reporting, the best test of reproducibility is to perform intercomparisons between laboratories, either by comparing results to identical experiments reported in the literature or, more robustly, through participation in planned intercomparisons (for example ‘round-robin’ exercises); we highlight a recent study applied to solid electrolyte characterisation as a topical example 43 . Intercomparisons are excellent at establishing the key features of an experimental method and the comparability of results obtained from different methods; moreover they provide a consensus against which other laboratories may compare themselves. However, performing repeat measurements for assessing repeatability and reproducibility cannot estimate uncertainty comprehensively, as it excludes systematic sources of uncertainty.

Assessing measurement uncertainty

Formal uncertainty evaluation is an alien concept to most electrochemists; even the best papers (as well as our own!) typically report only the standard deviation between a few repeats. Metrological best practice dictates that reported values are stated as the combination of a best estimate of the measurand, and an interval, and a coverage factor ( k ) which gives the probability of the true value being within that interval. For example, “the turnover frequency of the electrocatalyst is 1.0 ± 0.2 s −1 ( k  = 2)” 16 means that the scientist (having assumed normally distributed error) is 95% confident that the turnover frequency lies in the range 0.8–1.2 s −1 . Reporting results in such a way makes it immediately clear whether the measurements reliably support the stated conclusions, and enables meaningful comparisons between independent results even if their uncertainties differ (Fig.  3 ). It also encourages honesty and self-reflection about the shortcomings of results, encouraging the development of improved experimental techniques.

figure 3

a Complete reporting of a measurement includes the best estimate of the measurand and an uncertainty and the probability the true value falls within the uncertainty reported. Here, the percentages indicate that a normal distribution has been assumed. b Horizontal bars indicate 95% confidence intervals from uncertainty analysis. The confidence intervals of measurements 1 and 2 overlap when using k  = 2, so it is not possible to say with 95% confidence that the result of the measurement 2 is higher than measurement 1, but it is possible to say this with 68% confidence, i.e. k  = 1. Measurement 3 has a lower uncertainty, so it is possible to say with 95% confidence that the value is higher than measurement 2.

Constructing such a statement and performing the underlying calculations often appears daunting, not least as there are very few examples for electrochemical systems, with pH measurements being one example to have been treated thoroughly 44 . However, a standard process for uncertainty analysis exists, as briefly outlined graphically in Fig.  4 . We refer the interested reader to both accessible introductory texts 45 and detailed step-by-step guides 16 , 46 . The first steps in the process are to state precisely what is being measured—the measurand—and identify likely sources of uncertainty. Even this qualitative effort is often revealing. Precision in the definition of the measurand (and how it is determined from experimental indicators) clarifies the selection of measurement technique and helps to assess its appropriateness; for example, where the measurand relates only to an instantaneous property of a specific physical object, e.g. the current density of a specific fuel cell at 0.65 V following a standardised protocol, we ignore all variability in construction, device history etc. and no error is introduced by the sample. Whereas, when the measurand is a material property, such as turnover frequency of a catalyst material with a defined chemistry and preparation method, variability related to the material itself and sample preparation will often introduce substantial uncertainty in the final result. In electrochemical measurements, errors may arise from a range of sources including the measurement equipment, fluctuations in operating conditions, or variability in materials and samples. Identifying these sources leads to the design of better-quality experiments. In essence, the subsequent steps in the calculation of uncertainty quantify the uncertainty introduced by each source of error and, by using a measurement model or a sensitivity analysis (i.e. an assessment of how the results are sensitive to variability in input parameters), propagate these to arrive at a final uncertainty on the reported result.

figure 4

Possible sources of uncertainty are identified, and their standard uncertainty or probability distribution is determined by statistical analysis of repeat measurements (Type A uncertainties) or other evidence (Type B uncertainties). If required, uncertainties are then converted into the same unit as the measurand and adjusted for sensitivity, using a measurement model. Uncertainties are then combined either analytically using a standard approach or numerically to generate an overall estimate of uncertainty for the measurand (as indicated in Fig.  3a ).

Generally, given the historically poor understanding of uncertainty in electrochemistry, we promote increased awareness of uncertainty reporting standards and a focus on reporting measurement uncertainty with a level of detail that is appropriate to the claim made, or the scientific utilisation of the data. For example, where the primary conclusion of a paper relies on demonstrating that a material has the ‘highest ever X’ or ‘X is bigger than Y’ it is reasonable for reviewers to ask authors to quantify how confident they are in their measurement and statement. Additionally, where uncertainties are reported, even with error bars in numerical or graphical data, the method by which the uncertainty was determined should be stated, even if the method is consciously simple (e.g. “error bars indicate the sample standard deviation of n  = 3 measurements carried out on independent electrodes”). Unfortunately, we are aware of only sporadic and incomplete efforts to create formal uncertainty budgets for electrochemical measurements of energy technologies or materials, though work is underway in our group to construct these for some exemplar systems.

Electrochemistry has undoubtedly thrived without significant interaction with formal metrology; we do not urge an abrupt revolution whereby rigorous measurements become devalued if they lack additional arcane formality. Rather, we recommend using the well-honed principles of metrology to illuminate best practice and increase transparency about the strengths and shortcomings of reported experiments. From rethinking experimental design, to participating in laboratory intercomparisons and estimating the uncertainty on key results, the application of metrological principles to electrochemistry will result in more robust science.

Climent, V. & Feliu, J. M. Thirty years of platinum single crystal electrochemistry. J. Solid State Electrochem . https://doi.org/10.1007/s10008-011-1372-1 (2011).

Nature Editors and Contributors. Challenges in irreproducible research collection. Nature https://www.nature.com/collections/prbfkwmwvz/ (2018).

Chen, J. G., Jones, C. W., Linic, S. & Stamenkovic, V. R. Best practices in pursuit of topics in heterogeneous electrocatalysis. ACS Catal. 7 , 6392–6393 (2017).

Article   CAS   Google Scholar  

Voiry, D. et al. Best practices for reporting electrocatalytic performance of nanomaterials. ACS Nano 12 , 9635–9638 (2018).

Article   CAS   PubMed   Google Scholar  

Wei, C. et al. Recommended practices and benchmark activity for hydrogen and oxygen electrocatalysis in water splitting and fuel cells. Adv. Mater. 31 , 1806296 (2019).

Article   Google Scholar  

Chem Catalysis Editorial Team. Chem Catalysis Checklists Revision 1.1 . https://info.cell.com/introducing-our-checklists-learn-more (2021).

Chatenet, M. et al. Good practice guide for papers on fuel cells and electrolysis cells for the Journal of Power Sources. J. Power Sources 451 , 227635 (2020).

Sun, Y. K. An experimental checklist for reporting battery performances. ACS Energy Lett. 6 , 2187–2189 (2021).

Li, J. et al. Good practice guide for papers on batteries for the Journal of Power Sources. J. Power Sources 452 , 227824 (2020).

Arbizzani, C. et al. Good practice guide for papers on supercapacitors and related hybrid capacitors for the Journal of Power Sources. J. Power Sources 450 , 227636 (2020).

Hansen, J. N. et al. Is there anything better than Pt for HER? ACS Energy Lett. 6 , 1175–1180 (2021).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Xu, K. Navigating the minefield of battery literature. Commun. Mater. 3 , 1–7 (2022).

Dix, S. T., Lu, S. & Linic, S. Critical practices in rigorously assessing the inherent activity of nanoparticle electrocatalysts. ACS Catal. 10 , 10735–10741 (2020).

Mistry, A. et al. A minimal information set to enable verifiable theoretical battery research. ACS Energy Lett. 6 , 3831–3835 (2021).

Joint Committee for Guides in Metrology: Working Group 2. International Vocabulary of Metrology—Basic and General Concepts and Associated Terms . (2012).

Joint Committee for Guides in Metrology: Working Group 1. Evaluation of measurement data—Guide to the expression of uncertainty in measurement . (2008).

International Organization for Standardization: Committee on Conformity Assessment. ISO 17034:2016 General requirements for the competence of reference material producers . (2016).

Brown, R. J. C. & Andres, H. How should metrology bodies treat method-defined measurands? Accredit. Qual. Assur . https://doi.org/10.1007/s00769-020-01424-w (2020).

Angerstein-Kozlowska, H. Surfaces, Cells, and Solutions for Kinetics Studies . Comprehensive Treatise of Electrochemistry vol. 9: Electrodics: Experimental Techniques (Plenum Press, 1984).

Shinozaki, K., Zack, J. W., Richards, R. M., Pivovar, B. S. & Kocha, S. S. Oxygen reduction reaction measurements on platinum electrocatalysts utilizing rotating disk electrode technique. J. Electrochem. Soc. 162 , F1144–F1158 (2015).

International Organization for Standardization: Technical Committee ISO/TC 197. ISO 14687:2019(E)—Hydrogen fuel quality—Product specification . (2019).

Schmidt, T. J., Paulus, U. A., Gasteiger, H. A. & Behm, R. J. The oxygen reduction reaction on a Pt/carbon fuel cell catalyst in the presence of chloride anions. J. Electroanal. Chem. 508 , 41–47 (2001).

Chen, R. et al. Use of platinum as the counter electrode to study the activity of nonprecious metal catalysts for the hydrogen evolution reaction. ACS Energy Lett. 2 , 1070–1075 (2017).

Ji, S. G., Kim, H., Choi, H., Lee, S. & Choi, C. H. Overestimation of photoelectrochemical hydrogen evolution reactivity induced by noble metal impurities dissolved from counter/reference electrodes. ACS Catal. 10 , 3381–3389 (2020).

Jerkiewicz, G. Applicability of platinum as a counter-electrode material in electrocatalysis research. ACS Catal. 12 , 2661–2670 (2022).

Guo, J., Hsu, A., Chu, D. & Chen, R. Improving oxygen reduction reaction activities on carbon-supported ag nanoparticles in alkaline solutions. J. Phys. Chem. C. 114 , 4324–4330 (2010).

Arulmozhi, N., Esau, D., van Drunen, J. & Jerkiewicz, G. Design and development of instrumentations for the preparation of platinum single crystals for electrochemistry and electrocatalysis research Part 3: Final treatment, electrochemical measurements, and recommended laboratory practices. Electrocatal 9 , 113–123 (2017).

Colburn, A. W., Levey, K. J., O’Hare, D. & Macpherson, J. V. Lifting the lid on the potentiostat: a beginner’s guide to understanding electrochemical circuitry and practical operation. Phys. Chem. Chem. Phys. 23 , 8100–8117 (2021).

Ban, Z., Kätelhön, E. & Compton, R. G. Voltammetry of porous layers: staircase vs analog voltammetry. J. Electroanal. Chem. 776 , 25–33 (2016).

McMath, A. A., Van Drunen, J., Kim, J. & Jerkiewicz, G. Identification and analysis of electrochemical instrumentation limitations through the study of platinum surface oxide formation and reduction. Anal. Chem. 88 , 3136–3143 (2016).

Jerkiewicz, G. Standard and reversible hydrogen electrodes: theory, design, operation, and applications. ACS Catal. 10 , 8409–8417 (2020).

Ives, D. J. G. & Janz, G. J. Reference Electrodes, Theory and Practice (Academic Press, 1961).

Newman, J. & Balsara, N. P. Electrochemical Systems (Wiley, 2021).

Inzelt, G., Lewenstam, A. & Scholz, F. Handbook of Reference Electrodes (Springer Berlin, 2013).

Cimenti, M., Co, A. C., Birss, V. I. & Hill, J. M. Distortions in electrochemical impedance spectroscopy measurements using 3-electrode methods in SOFC. I—effect of cell geometry. Fuel Cells 7 , 364–376 (2007).

Hoshi, Y. et al. Optimization of reference electrode position in a three-electrode cell for impedance measurements in lithium-ion rechargeable battery by finite element method. J. Power Sources 288 , 168–175 (2015).

Article   ADS   CAS   Google Scholar  

Jackson, C., Lin, X., Levecque, P. B. J. & Kucernak, A. R. J. Toward understanding the utilization of oxygen reduction electrocatalysts under high mass transport conditions and high overpotentials. ACS Catal. 12 , 200–211 (2022).

Masa, J., Batchelor-McAuley, C., Schuhmann, W. & Compton, R. G. Koutecky-Levich analysis applied to nanoparticle modified rotating disk electrodes: electrocatalysis or misinterpretation. Nano Res. 7 , 71–78 (2014).

Martens, S. et al. A comparison of rotating disc electrode, floating electrode technique and membrane electrode assembly measurements for catalyst testing. J. Power Sources 392 , 274–284 (2018).

Wei, C. et al. Approaches for measuring the surface areas of metal oxide electrocatalysts for determining their intrinsic electrocatalytic activity. Chem. Soc. Rev. 48 , 2518–2534 (2019).

Lukaszewski, M., Soszko, M. & Czerwiński, A. Electrochemical methods of real surface area determination of noble metal electrodes—an overview. Int. J. Electrochem. Sci. https://doi.org/10.20964/2016.06.71 (2016).

Niu, S., Li, S., Du, Y., Han, X. & Xu, P. How to reliably report the overpotential of an electrocatalyst. ACS Energy Lett. 5 , 1083–1087 (2020).

Ohno, S. et al. How certain are the reported ionic conductivities of thiophosphate-based solid electrolytes? An interlaboratory study. ACS Energy Lett. 5 , 910–915 (2020).

Buck, R. P. et al. Measurement of pH. Definition, standards, and procedures (IUPAC Recommendations 2002). Pure Appl. Chem. https://doi.org/10.1351/pac200274112169 (2003).

Bell, S. Good Practice Guide No. 11. The Beginner’s Guide to Uncertainty of Measurement. (Issue 2). (National Physical Laboratory, 2001).

United Kingdom Accreditation Service. M3003 The Expression of Uncertainty and Confidence in Measurement  4th edn. (United Kingdom Accreditation Service, 2019).

Download references

Acknowledgements

This work was supported by the National Measurement System of the UK Department of Business, Energy & Industrial Strategy. Andy Wain, Richard Brown and Gareth Hinds (National Physical Laboratory, Teddington, UK) provided insightful comments on the text.

Author information

Authors and affiliations.

National Physical Laboratory, Hampton Road, Teddington, TW11 0LW, UK

Graham Smith & Edmund J. F. Dickinson

You can also search for this author in PubMed   Google Scholar

Contributions

G.S. originated the concept for the article. G.S. and E.J.F.D. contributed to drafting, editing and revision of the manuscript.

Corresponding author

Correspondence to Graham Smith .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Peer review

Peer review information.

Nature Communications thanks Gregory Jerkiewicz for their contribution to the peer review of this work.

Additional information

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Smith, G., Dickinson, E.J.F. Error, reproducibility and uncertainty in experiments for electrochemical energy technologies. Nat Commun 13 , 6832 (2022). https://doi.org/10.1038/s41467-022-34594-x

Download citation

Received : 29 July 2022

Accepted : 31 October 2022

Published : 11 November 2022

DOI : https://doi.org/10.1038/s41467-022-34594-x

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Electrochemical surface-enhanced raman spectroscopy.

  • Christa L. Brosseau
  • Alvaro Colina

Nature Reviews Methods Primers (2023)

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

short definition of experimental error

Incremental Plastic Analysis of Confined Concrete Considering the Variation of Elastic Moduli

  • Research Paper
  • Published: 16 September 2024

Cite this article

short definition of experimental error

  • Mohammad Rasouli 1 ,
  • Saeed Baghdarnia 1 &
  • Vahid Broujerdian   ORCID: orcid.org/0000-0003-3454-4797 1  

Experimental studies have demonstrated that the elastic moduli of concrete, specifically Young’s modulus and Poisson’s ratio, undergo changes during compressive loading. Despite the fact that variations in Young’s modulus are frequently considered in nonlinear analyses, Poisson’s ratio is typically assumed to be constant, which has a direct impact on confinement modeling. In this research project, an attempt was made to enhance the accuracy of predicting the behavior of concrete columns confined by AFRP and CFRP by considering the variation of elastic moduli of concrete during loading. To account for the changes of Poisson’s ratio, an approximate method was proposed that involves assembling a three-part stress–strain curve. The first and last parts of the curve coincide with the stress–strain curves obtained by the limit Poisson’s ratio of 0.2 and 0.5, respectively, while a linear function serves as the transition curve in the middle region. The parameters of the middle zone were calculated using two different approaches: the first involved data fitting and optimization, while the second entailed using a proposed closed-form equation. The finite element program ABAQUS was employed to conduct incremental plastic analyses within the Concrete Damage Plasticity framework. The proposed model is capable of predicting the complete axial compressive stress–strain curve of concrete columns confined by AFRP and CFRP under monotonic compressive loading. A corroboration study was conducted using an experimental dataset from 24 concrete short column test specimens confined by AFRP and CFRP with a wide range of properties. The results showed that the average errors of both the proposed methods are nearly 3%. It means that both the numerical methods generally have a similar and acceptable precision.

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

Access this article

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

short definition of experimental error

Data availability

Data will be made available on request.

Abbreviations

Average absolute error

Aramid fiber reinforced polymer

Concrete damage plasticity

Carbon fiber reinforced polymer

Coefficient of variation

The C3D8R element is a general-purpose linear brick element, with reduced integration (1 integration point)

R-squared error

Standard deviation

S4R is a 4-node, quadrilateral, stress/displacement shell element with reduced integration

Abbasnia R, Holakoo A (2012) An investigation of stress-strain behavior of FRP-confined concrete under cyclic compressive loading. Int J Civil Eng 10(3):201–209

Google Scholar  

Bai YL, Yang HL, Mohammadi M, Dai JG (2023) 3D FE analysis for buckling behavior of longitudinal reinforcement embedded in LRS FRP-confined concrete under axial compression. Compos Struct 312:116857. https://doi.org/10.1016/j.compstruct.2023.116857

Article   Google Scholar  

Benmokrane B, Xu H, Nishizaki I (1997) Aramid and carbon fibre-reinforced plastic prestressed ground anchors and their field applications. Can J Civ Eng 24(6):968–985. https://doi.org/10.1139/cjce-24-6-968

Carreira DJ, Chu K-H (1985) Stress-strain relationship for plain concrete in compression. ACI J Proceed 82(6):797. https://doi.org/10.14359/10390

Chaallal O, Hassan M, Shahawy M (2003) Confinement model for axially loaded short rectangular columns strengthened with fiber-reinforced polymer wrapping. ACI Struct J 100(2):215–221. https://doi.org/10.14359/12485

Chen B, Wang L (2019) Experimental study on hollow steel-reinforced concrete-filled GFRP tubular members under axial compression. Steel Compos Struct 32(1):59–66. https://doi.org/10.12989/scs.2019.32.1.059

Chen WF (2007) Plasticity in reinforced concrete: J. Ross Publishing

Devore JL (2015) Probability and statistics for engineering and the sciences: Cengage learning

Djenad S, Taleb SA, Salem AS (2023) Finite element modeling of partially-confined concrete and RC columns with embedded hexagonal-FRP strips under axial and horizontal loading. Structures 54:369–385. https://doi.org/10.1016/j.istruc.2023.05.065

Elwi AA, Murray DW (1979) A 3D hypoelastic concrete constitutive relationship. J Eng Mech Div 105(4):623–641. https://doi.org/10.1061/JMCEA3.0002510

Hassan GB, Al-Kamaki YSS, Mohammed AA, AlSaad A (2022) Long-term exposure of RC columns immersed in seawater or crude oil confined with CFRP fabrics under monotonic or cyclic loading. Case Stud Constr Mater 18:e01747. https://doi.org/10.1016/j.cscm.2022.e01747

Hibbett K, Sorensen (1998) ABAQUS/standard: user’s manual (Vol. 1): Hibbitt Karlsson and Sorensen

Hintze W, Hartmann D (2013) Modeling of delamination during milling of unidirectional CFRP. Procedia Cirp 8:444–449

Hsu L, Hsu C-T (1994) Complete stress—strain behaviour of high-strength concrete under compression. Mag Concr Res 46(169):301–312. https://doi.org/10.1680/macr.1994.46.169.301

Isleem HF, Tahir M, Wang Z (2020) Axial stress–strain model developed for rectangular RC columns confined with FRP wraps and anchors. Structures 23:779–788. https://doi.org/10.1016/j.istruc.2019.12.020

Isleem HF, Zewudie BB, Bahrami A, Kumar R, Xingchong W, Samui P (2024) Parametric investigation of rectangular CFRP-confined concrete columns reinforced by inner elliptical steel tubes using finite element and machine learning models. Heliyon 10(2):e23666. https://doi.org/10.1016/j.heliyon.2023.e23666

Issa CA, Chami P, Saad G (2009) Compressive strength of concrete cylinders with variable widths CFRP wraps: experimental study and numerical modeling. Constr Build Mater 23(6):2306–2318. https://doi.org/10.1016/j.conbuildmat.2008.11.009

Karabinis AI, Rousakis TC, Manolitsi GE (2008) 3D finite-element analysis of substandard RC columns strengthened by fiber-reinforced polymer sheets. J Compos Constr 12(5):531–540. https://doi.org/10.1061/(ASCE)1090-0268(2008)12:5(531)

Koksal H, Doran B, Turgay T (2009) A practical approach for modeling FRP wrapped concrete columns. Constr Build Mater 23(3):1429–1437. https://doi.org/10.1016/j.conbuildmat.2008.07.008

Kupfer H, Hilsdorf HK, Rusch H (1969) Behavior of concrete under biaxial stresses. ACI J Proceed 66(8):656. https://doi.org/10.14359/7388

Lam L, Teng J (2003a) Design-oriented stress-strain model for FRP-confined concrete in rectangular columns. J Reinf Plast Compos 22(13):1149–1186. https://doi.org/10.1177/0731684403035429

Lam L, Teng J (2003b) Design-oriented stress–strain model for FRP-confined concrete. Constr Build Mater 17(6–7):471–489

Li P, Wu Y-F (2016) Stress–strain behavior of actively and passively confined concrete under cyclic axial load. Compos Struct 149:369–384. https://doi.org/10.1016/j.compstruct.2016.04.033

Lin G, Teng J (2017) Three-dimensional finite-element analysis of FRP-confined circular concrete columns under eccentric loading. J Compos Constr 21(4):04017003. https://doi.org/10.1061/(ASCE)CC.1943-5614.0000772

Lokuge W, Otoom O, Borzou R, Navaratnam S, Herath N, Thambiratnam D (2021) Experimental and numerical analysis on the effectiveness of GFRP wrapping system on timber pile rehabilitation. Case Stud Constr Mater 15:e00552. https://doi.org/10.1016/j.cscm.2021.e00552

Loo Y, Base G (1990) Variation of creep Poisson’s ratio with stress in concrete under short-term uniaxial compression. Mag Concr Res 42(151):67–73. https://doi.org/10.1680/macr.1990.42.151.67

Lubliner J, Oliver J, Oller S, Oñate E (1989) A plastic-damage model for concrete. Int J Solids Struct 25(3):299–326. https://doi.org/10.1016/0020-7683(89)90050-4

Maheri MR, Torabi A (2019) Retrofitting external RC beam-column joints of an ordinary MRF through plastic hinge relocation using FRP laminates. Pap Present Struct. https://doi.org/10.1016/j.istruc.2019.08.004

Mirmiran A, Shahawy M (1997) Dilation characteristics of confined concrete. Mech Cohes Frict Mater Int J Exp Modell Comput Mater Struct 2(3):237–249

Mirmiran A, Zagers K, Yuan W (2000) Nonlinear finite element modeling of concrete confined by fiber composites. Finite Elem Anal Des 35(1):79–96. https://doi.org/10.1016/S0168-874X(99)00056-6

Otoom OF, Lokuge W, Karunasena W, Manalo AC, Ozbakkaloglu T, Thambiratnam D (2021) Experimental and numerical evaluation of the compression behaviour of GFRP-wrapped infill materials. Case Stud Constr Mater 15:e00654. https://doi.org/10.1016/j.cscm.2021.e00654

Ottosen NS (1979) Constitutive model for short-time loading of concrete. J Eng Mech Div ASCE 105:127–141. https://doi.org/10.1061/JMCEA3.0002446

Ozbakkaloglu T, Lim JC, Vincent T (2013) FRP-confined concrete in circular sections: review and assessment of stress–strain models. Eng Struct 49:1068–1088. https://doi.org/10.1016/j.engstruct.2012.06.010

Rasouli M, Broujerdian V (2019) 3D finite element modeling of FRP-confined rectangular short columns considering variation of Poisson’s ratio. Iran J Sci Technol Trans Civil Eng. https://doi.org/10.1007/s40996-019-00276-w

Rasouli M, Broujerdian V, Kazemnadi A (2020) Predicting the compressive stress–strain curve of FRP-confined concrete column considering the variation of Poisson’s ratio. Int J Civil Eng 18:1365–1380. https://doi.org/10.1007/s40999-020-00550-3

Rodsin K, Hussain Q, Suparp S, Nawaz A (2020) Compressive behavior of extremely low strength concrete confined with low-cost glass FRP composites. Case Stud Constr Mater 13:e00452. https://doi.org/10.1109/ESIT.2018.8665331

Samaan M, Mirmiran A, Shahawy M (1998) Model of concrete confined by fiber composites. J Struct Eng 124(9):1025–1031

Selby RG, Vecchio FJ (1997) A constitutive model for analysis of reinforced concrete solids. Can J Civ Eng 24(3):460–470. https://doi.org/10.1139/l96-135

Shahawy M, Mirmiran A, Beitelman T (2000) Tests and modeling of carbon-wrapped concrete columns. Compos B Eng 31(6–7):471–480. https://doi.org/10.1016/S1359-8368(00)00021-4

Spoelstra MR, Monti G (1999) FRP-confined concrete model. J Compos Constr 3(3):143–150. https://doi.org/10.1061/(ASCE)1090-0268(1999)3:3(143)

Vuggumudi S, Alagusundaramoorthy P (2018) FRP strengthened RC rectangular columns under combined axial and lateral loading: analytical study. Paper Present Struct. https://doi.org/10.1016/j.istruc.2018.02.007

Wang Y-F, Wu H-L (2011) Size effect of concrete short columns confined with aramid FRP jackets. J Compos Constr 15(4):535–544. https://doi.org/10.1061/(ASCE)CC.1943-5614.0000178

Wang J, Liu W, Zhou D, Zhu L, Fang H (2014) Mechanical behaviour of concrete filled double skin steel tubular stub columns confined by FRP under axial compression. Steel Compos Struct 17(4):431–452. https://doi.org/10.12989/scs.2014.17.4.431

Wu G, Lü Z, Wu Z (2006) Strength and ductility of concrete cylinders confined with FRP composites. Constr Build Mater 20(3):134–148. https://doi.org/10.1016/j.conbuildmat.2005.01.022

Xia S, Wei X, He C, Ding F, Zhang W, Wu J, Yang S (2022) Performance of stirrup-confined concrete-filled steel tubular stub columns under axial loading: a further investigation. Case Stud Constr Mater 17:e01483. https://doi.org/10.1016/j.cscm.2022.e01483

Xiao Y, Wu H (2000) Compressive behavior of concrete confined by carbon fiber composite jackets. J Mater Civ Eng 12(2):139–146. https://doi.org/10.1061/(ASCE)0899-1561(2000)12:2(139)

Yu T, Teng J, Wong Y, Dong S (2010) Finite element modeling of confined concrete-II: plastic-damage model. Eng Struct 32(3):680–691. https://doi.org/10.1016/j.engstruct.2009.11.013

Yu F, Zhang N, Fang Y, Liu J, Xiang G (2020) Stress-strain model of weak PVC-FRP confined concrete column and strong RC ring beam joint under eccentric compression. Steel Compos Struct Int J 35(1):13–27. https://doi.org/10.12989/scs.2020.35.1.013

Zhao L, Li Y, Zhang J, Zhou L, Hu N (2018) A novel material degradation model for unidirectional CFRP composites. Compos B Eng 135:84–94. https://doi.org/10.1016/j.compositesb.2017.09.038

Download references

Author information

Authors and affiliations.

School of Civil Engineering, Iran University of Science and Technology, Tehran, Iran

Mohammad Rasouli, Saeed Baghdarnia & Vahid Broujerdian

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Vahid Broujerdian .

Ethics declarations

Conflict of interest.

The authors declare that they have no conflict of interest.

Additional information

Publisher's note.

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

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Rasouli, M., Baghdarnia, S. & Broujerdian, V. Incremental Plastic Analysis of Confined Concrete Considering the Variation of Elastic Moduli. Iran J Sci Technol Trans Civ Eng (2024). https://doi.org/10.1007/s40996-024-01613-4

Download citation

Received : 28 August 2023

Accepted : 03 September 2024

Published : 16 September 2024

DOI : https://doi.org/10.1007/s40996-024-01613-4

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Confinement
  • Finite element method
  • Poisson’s ratio
  • Find a journal
  • Publish with us
  • Track your research

IMAGES

  1. PPT

    short definition of experimental error

  2. PPT

    short definition of experimental error

  3. PPT

    short definition of experimental error

  4. PPT

    short definition of experimental error

  5. Error Analysis

    short definition of experimental error

  6. PPT

    short definition of experimental error

VIDEO

  1. Definition of Experimental Error III BSc Stat P5 U5 L4

  2. Experimental methods and Error analysis

  3. Michelson's interferometer

  4. Building a Compiler

  5. Adriano Celentano's Music Video Experiment

  6. Concert of Within Experimental Error 2.0 at ICMM, June 20 2024

COMMENTS

  1. Experimental Error Types, Sources & Examples

    As a member, you'll also get unlimited access to over 88,000 lessons in math, English, science, history, and more. Plus, get practice tests, quizzes, and personalized coaching to help you succeed.

  2. Understanding Experimental Errors: Types, Causes, and Solutions

    These errors are often classified into three main categories: systematic errors, random errors, and human errors. Here are some common types of experimental errors: 1. Systematic Errors. Systematic errors are consistent and predictable errors that occur throughout an experiment. They can arise from flaws in equipment, calibration issues, or ...

  3. PDF Introduction to Error and Uncertainty

    from experimental data. In this lab course, we will be using Microsoft Excel to record ... Systematic errors are usually due to imperfections in the equipment, improper or biased observation, or the presence of additional physical e ects not taken into account. (An example might be an experiment on forces and acceleration in which

  4. PDF Chapter 3 Experimental Error

    Consider the function pH = −log [H+], where [H+] is the molarity of H+. For pH = 5.21 ± 0.03, find [H+] and its uncertainty. The concentration of H+ is 6.17 (±0.426) × 10−6 = 6.2 (±0.4) × l0−6 M. The number of significant digits in a number is the required to write the number in scientific notation.

  5. Sources of Error in Science Experiments

    Random errors are due to fluctuations in the experimental or measurement conditions. Usually these errors are small. Taking more data tends to reduce the effect of random errors. Examples of Random Errors

  6. PDF An Introduction to Experimental Uncertainties and Error Analysis

    Lynn 4 Uncertainties and Error Analysis What Is an Error Bar? In a laboratory setting, or in any original, quantitative research, we make our

  7. Random vs. Systematic Error

    Random and systematic errors are types of measurement error, a difference between the observed and true values of something.

  8. PDF Experimental Uncertainties (Errors)

    There are three main sources of experimental uncertainties (experimental errors): 1. Limited accuracy of the measuring apparatus - e.g., the force sensors that we use in experiment M2 cannot determine applied force with a better accuracy than ±0.05 N. 2. Limitations and simplifications of the experimental procedure - e.g., we commonly

  9. PDF ERROR ANALYSIS (UNCERTAINTY ANALYSIS)

    4 USES OF UNCERTAINTY ANALYSIS (I) • Assess experimental procedure including identification of potential difficulties - Definition of necessary steps - Gaps • Advise what procedures need to be put in place for measurement • Identify instruments and procedures that control accuracy and precision - Usually one, or at most a small number, out of the large set of

  10. PDF Notes on Error Analysis

    Systematic errors usually shift measurements in a systematic way. They can be built into instruments. Systematic errors can be at least minimized by instrument calibration and appropriate use of equipment. Extraneous effects can also alter experimental results. The terms accuracy and precision are often misused. Experimental precision means the

  11. Types of Error

    Random errors occur randomly, and sometimes have no source/cause. There are two types of random errors. Observational: When the observer makes consistent observational mistakes (such not reading the scale correctly and writing down values that are constantly too low or too high) Environmental: When unpredictable changes occur in the environment ...

  12. Experimental Errors and Error Analysis

    This chapter is largely a tutorial on handling experimental errors of measurement. Much of the material has been extensively tested with science undergraduates at a variety of levels at the University of Toronto. ... Calibration standards are, almost by definition, too delicate and/or expensive to use for direct measurement. Here is an example ...

  13. PDF A Student's Guide to Data and Error Analysis

    Preface. This book is written as a guide for the presentation of experimental including a consistent treatment of experimental errors and inaccuracies. is meant for experimentalists in physics, astronomy, chemistry, life and engineering. However, it can be equally useful for theoreticians produce simulation data: they are often confronted with ...

  14. Error, reproducibility and uncertainty in experiments for

    Precision in the definition of the measurand (and how it is determined from experimental indicators) clarifies the selection of measurement technique and helps to assess its appropriateness; for ...

  15. Incremental Plastic Analysis of Confined Concrete Considering the

    A corroboration study was conducted using an experimental dataset from 24 concrete short column test specimens confined by AFRP and CFRP with a wide range of properties. The results showed that the average errors of both the proposed methods are nearly 3%. It means that both the numerical methods generally have a similar and acceptable precision.