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.

Receive feedback on language, structure, and formatting

Professional editors proofread and edit your paper by focusing on:

  • Academic style
  • Vague sentences
  • Style consistency

See an example

types of errors in a science experiment

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 August 12, 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.

Learn The Types

Learn About Different Types of Things and Unleash Your Curiosity

Understanding Experimental Errors: Types, Causes, and Solutions

Types of experimental errors.

In scientific experiments, errors can occur that affect the accuracy and reliability of the results. 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 flawed experimental design. Some examples of systematic errors include:

– Instrumental Errors: These errors occur due to inaccuracies or limitations of the measuring instruments used in the experiment. For example, a thermometer may consistently read temperatures slightly higher or lower than the actual value.

– Environmental Errors: Changes in environmental conditions, such as temperature or humidity, can introduce systematic errors. For instance, if an experiment requires precise temperature control, fluctuations in the room temperature can impact the results.

– Procedural Errors: Errors in following the experimental procedure can lead to systematic errors. This can include improper mixing of reagents, incorrect timing, or using the wrong formula or equation.

2. Random Errors

Random errors are unpredictable variations that occur during an experiment. They can arise from factors such as inherent limitations of measurement tools, natural fluctuations in data, or human variability. Random errors can occur independently in each measurement and can cause data points to scatter around the true value. Some examples of random errors include:

– Instrument Noise: Instruments may introduce random noise into the measurements, resulting in small variations in the recorded data.

– Biological Variability: In experiments involving living organisms, natural biological variability can contribute to random errors. For example, in studies involving human subjects, individual differences in response to a treatment can introduce variability.

– Reading Errors: When taking measurements, human observers can introduce random errors due to imprecise readings or misinterpretation of data.

3. Human Errors

Human errors are mistakes or inaccuracies that occur due to human factors, such as lack of attention, improper technique, or inadequate training. These errors can significantly impact the experimental results. Some examples of human errors include:

– Data Entry Errors: Mistakes made when recording data or entering data into a computer can introduce errors. These errors can occur due to typographical mistakes, transposition errors, or misinterpretation of results.

– Calculation Errors: Errors in mathematical calculations can occur during data analysis or when performing calculations required for the experiment. These errors can result from mathematical mistakes, incorrect formulas, or rounding errors.

– Experimental Bias: Personal biases or preconceived notions held by the experimenter can introduce bias into the experiment, leading to inaccurate results.

It is crucial for scientists to be aware of these types of errors and take measures to minimize their impact on experimental outcomes. This includes careful experimental design, proper calibration of instruments, multiple repetitions of measurements, and thorough documentation of procedures and observations.

You Might Also Like:

Patio perfection: choosing the best types of pavers for your outdoor space, a guide to types of pupusas: delicious treats from central america, exploring modern period music: from classical to jazz and beyond.

Types of Error — Overview & Comparison - Expii

  • Faculty Resource Center
  • Biochemistry
  • Bioengineering
  • Cancer Research
  • Developmental Biology
  • Engineering
  • Environment
  • Immunology and Infection
  • Neuroscience
  • JoVE Journal
  • JoVE Encyclopedia of Experiments
  • JoVE Chrome Extension
  • Environmental Sciences
  • Pharmacology
  • JoVE Science Education
  • JoVE Lab Manual
  • JoVE Business
  • Videos Mapped to your Course
  • High Schools
  • Videos Mapped to Your Course

Chapter 1: Chemical Applications of Statistical Analyses

Back to chapter, types of errors: detection and minimization, previous video 1.2: degrees of freedom, next video 1.5: systematic error: methodological and sampling errors.

Error is the deflection of an obtained result from the expected or true results of an experiment. This happens due to the uncertainty associated with the measurement.

Errors can be expressed in absolute or relative terms.

While absolute error is the difference between the measured and the true value, the relative error is the percentage of absolute error to the actual value.

Errors can be classified based on the source, magnitude, and sign into systematic, random, and gross errors.

Systematic errors are reproducible as they originate from defective equipment, flawed experiment design, and personal bias.

Random errors arise from uncontrollable variables in the measurement, making them irreproducible and randomly scattered around a central value.

Gross errors occur due to human mistakes and are of larger magnitude.

Systematic errors can be detected and minimized by using standard reference materials, independent analysis, blank determinations, and varying the sample size.

Error is the deviation of the obtained result from the true, expected value or the estimated central value. Errors are expressed in absolute or relative terms.

Absolute error in a measurement is the numerical difference from the true or central value. Relative error is the ratio between absolute error and the true or central value, expressed as a percentage.

Errors can be classified by source, magnitude, and sign. There are three types of errors: systematic, random, and gross.

Systematic or determinate errors emerge from known sources and are reproducible during replicate measurements. Defective equipment and experiment design flaws are familiar sources of these errors. These errors can be minimized by employing standard reference materials, independent analysis, or varying the sample size.

Random errors or indeterminate errors are difficult to reproduce by repeating measurements. These errors originate from uncontrolled variables like electronic noise in the circuit of an electrical instrument and irregular changes in the heat loss rate from a solar collector due to changes in the wind.

Gross errors are caused by human mistakes. The magnitude of these errors is often high. The origin of such errors is entirely based on the observer.

Simple Hit Counter

Back Home

  • Science Notes Posts
  • Contact Science Notes
  • Todd Helmenstine Biography
  • Anne Helmenstine Biography
  • Free Printable Periodic Tables (PDF and PNG)
  • Periodic Table Wallpapers
  • Interactive Periodic Table
  • Periodic Table Posters
  • Science Experiments for Kids
  • How to Grow Crystals
  • Chemistry Projects
  • Fire and Flames Projects
  • Holiday Science
  • Chemistry Problems With Answers
  • Physics Problems
  • Unit Conversion Example Problems
  • Chemistry Worksheets
  • Biology Worksheets
  • Periodic Table Worksheets
  • Physical Science Worksheets
  • Science Lab Worksheets
  • My Amazon Books

Systematic vs Random Error – Differences and Examples

Systematic Error vs Random Error

Systematic and random error are an inevitable part of measurement. Error is not an accident or mistake. It naturally results from the instruments we use, the way we use them, and factors outside our control. Take a look at what systematic and random error are, get examples, and learn how to minimize their effects on measurements.

  • Systematic error has the same value or proportion for every measurement, while random error fluctuates unpredictably.
  • Systematic error primarily reduces measurement accuracy, while random error reduces measurement precision.
  • It’s possible to reduce systematic error, but random error cannot be eliminated.

Systematic vs Random Error

Systematic error is consistent, reproducible error that is not determined by chance. Systematic error introduces inaccuracy into measurements, even though they may be precise. Averaging repeated measurements does not reduce systematic error, but calibrating instruments helps. Systematic error always occurs and has the same value when repeating measurements the same way.

As its name suggests, random error is inconsistent error caused by chance differences that occur when taking repeated measurements. Random error reduces measurement precision, but measurements cluster around the true value. Averaging measurements containing only random error gives an accurate, imprecise value. Random errors cannot be controlled and are not the same from one measurement to the next.

Systematic Error Examples and Causes

Systematic error is consistent or proportional to the measurement, so it primarily affects accuracy. Causes of systematic error include poor instrument calibration, environmental influence, and imperfect measurement technique.

Here are examples of systematic error:

  • Reading a meniscus above or below eye level always gives an inaccurate reading. The reading is consistently high or low, depending on the viewing angle.
  • A scale gives a mass measurement that is always “off” by a set amount. This is called an offset error . Taring or zeroing a scale counteracts this error.
  • Metal rulers consistently give different measurements when they are cold compared to when they are hot due to thermal expansion. Reducing this error means using a ruler at the temperature at which it was calibrated.
  • An improperly calibrated thermometer gives accurate readings within a normal temperature range. But, readings become less accurate at higher or lower temperatures.
  • An old, stretched cloth measuring tape gives consistent, but different measurements than a new tape. Proportional errors of this type are called scale factor errors .
  • Drift occurs when successive measurements become consistently higher or lower as time progresses. Electronic equipment is susceptible to drift. Devices that warm up tend to experience positive drift. In some cases, the solution is to wait until an instrument warms up before using it. In other cases, it’s important to calibrate equipment to account for drift.

How to Reduce Systematic Error

Once you recognize systematic error, it’s possible to reduce it. This involves calibrating equipment, warming up instruments because taking readings, comparing values against standards, and using experimental controls. You’ll get less systematic error if you have experience with a measuring instrument and know its limitations. Randomizing sampling methods also helps, particularly when drift is a concern.

Random Error Examples and Causes

Random error causes measurements to cluster around the true value, so it primarily affects precision. Causes of random error include instrument limitations, minor variations in measuring techniques, and environmental factors.

Here are examples of random error:

  • Posture changes affect height measurements.
  • Reaction speed affects timing measurements.
  • Slight variations in viewing angle affect volume measurements.
  • Wind velocity and direction measurements naturally vary according to the time at which they are taken. Averaging several measurements gives a more accurate value.
  • Readings that fall between the marks on a device must be estimated. To some extent, its possible to minimize this error by choosing an appropriate instrument. For example, volume measurements are more precise using a graduated cylinder instead of a beaker.
  • Mass measurements on an analytical balance vary with air currents and tiny mass changes in the sample.
  • Weight measurements on a scale vary because it’s impossible to stand on the scale exactly the same way each time. Averaging multiple measurements minimizes the error.

How to Reduce Random Error

It’s not possible to eliminate random error, but there are ways to minimize its effect. Repeat measurements or increase sample size. Be sure to average data to offset the influence of chance.

Which Types of Error Is Worse?

Systematic errors are a bigger problem than random errors. This is because random errors affect precision, but it’s possible to average multiple measurements to get an accurate value. In contrast, systematic errors affect precision. Unless the error is recognized, measurements with systematic errors may be far from true values.

  • Bland, J. Martin, and Douglas G. Altman (1996). “Statistics Notes: Measurement Error.”  BMJ  313.7059: 744.
  • Cochran, W. G. (1968). “Errors of Measurement in Statistics”.  Technometrics . Taylor & Francis, Ltd. on behalf of American Statistical Association and American Society for Quality. 10: 637–666. doi: 10.2307/1267450
  • Dodge, Y. (2003).  The Oxford Dictionary of Statistical Terms . OUP. ISBN 0-19-920613-9.
  • Taylor, J. R. (1999).  An Introduction to Error Analysis: The Study of Uncertainties in Physical Measurements . University Science Books. ISBN 0-935702-75-X.

Related Posts

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

types of errors in a science experiment

  • 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
  • Units Of Measurement
  • Accuracy Precision Error Measurement

Errors in Measurement: Gross Errors, Systematic Errors and Random Errors

Measurements are an integral part of living; we measure time, measure steps walked to know the calories burnt, measure the materials added for cooking, and measure the size of clothes to know whether it fits perfectly. Sometimes we fail to know the exact measurement, and the values vary, leading to errors. In this article, let us learn about measurement, errors in measurement, types of errors and how to avoid the errors.

Table of Contents:

Measurement

Types of errors, errors calculation, how to reduce errors in measurement.

  • Frequently Asked Questions – FAQs

Measurement is the foundation for all experimental science. All the great technological development could not have been possible without ever-increasing levels of accuracy of measurements . The measurement of an amount is based on some international standards, which are completely accurate compared with others. Just like your vegetable vendors, measurements are taken by comparing an unknown amount with a known weight. Every measurement carries a level of uncertainty which is known as an error. This error may arise in the process or due to a mistake in the experiment. So 100% accurate measurement is not possible with any method.

An error may be defined as the difference between the measured and actual values. For example, if the two operators use the same device or instrument for measurement. It is not necessary that both operators get similar results. The difference between the measurements is referred to as an ERROR.

To understand the concept of measurement errors, you should know the two terms that define the error. They are true value and measured value. The true value is impossible to find by experimental means. It may be defined as the average value of an infinite number of measured values. The measured value is a single measure of the object to be as accurate as possible.

There are three types of errors that are classified based on the source they arise from; They are:

  • Gross Errors
  • Random Errors
  • Systematic Errors

This category basically takes into account human oversight and other mistakes while reading, recording, and readings. The most common human error in measurement falls under this category of measurement errors. For example, the person taking the reading from the meter of the instrument may read 23 as 28. Gross errors can be avoided by using two suitable measures, and they are written below:

  • Proper care should be taken in reading, recording the data. Also, the calculation of error should be done accurately.
  • By increasing the number of experimenters, we can reduce the gross errors. If each experimenter takes different readings at different points, then by taking the average of more readings, we can reduce the gross errors

The random errors are those errors, which occur irregularly and hence are random. These can arise due to random and unpredictable fluctuations in experimental conditions ( Example : unpredictable fluctuations in temperature, voltage supply, mechanical vibrations of experimental set-ups, etc, errors by the observer taking readings, etc. For example, when the same person repeats the same observation, he may likely get different readings every time.

This article explored the various types of errors in the measurements we make. These errors are everywhere in every measurement we make. To find more articles, visit BYJU’S. Join us and fall in love with learning.

Systematic Errors:

Systematic errors can be better understood if we divide them into subgroups; They are:

  • Environmental Errors
  • Observational Errors
  • Instrumental Errors

Environmental Errors: This type of error arises in the measurement due to the effect of the external conditions on the measurement. The external condition includes temperature, pressure, and humidity and can also include an external magnetic field . If you measure your temperature under the armpits and during the measurement, if the electricity goes out and the room gets hot, it will affect your body temperature, affecting the reading.

Observational Errors: These are the errors that arise due to an individual’s bias, lack of proper setting of the apparatus, or an individual’s carelessness in taking observations. The measurement errors also include wrong readings due to Parallax errors.

Instrumental Errors: These errors arise due to faulty construction and calibration of the measuring instruments. Such errors arise due to the hysteresis of the equipment or due to friction . Lots of the time, the equipment being used is faulty due to misuse or neglect, which changes the reading of the equipment. The zero error is a very common type of error. This error is common in devices like Vernier callipers and screw gauges. The zero error can be either positive or negative. Sometimes the scale readings are worn off, which can also lead to a bad reading.

Instrumental error takes place due to :

  • An inherent constraint of devices 
  • Misuse of Apparatus 
  • Effect of Loading

Different measures of errors include:

Absolute Error

The difference between the measured value of a quantity and its actual value gives the absolute error. It is the variation between the actual values and measured values. It is given by 

Absolute error = |VA-VE|

Percent Error

It is another way of expressing the error in measurement. This calculation allows us to gauge how accurate a measured value is with respect to the true value. Per cent error is given by the formula

Percentage error (%) = (VA-VE) / VE) x 100

Relative Error

The ratio of the absolute error to the accepted measurement gives the relative error. The relative error is given by the formula:

Relative Error = Absolute error / Actual value

Keeping an eye on the procedure and following the below listed points can help to reduce the error.

  • Make sure the formulas used for measurement are correct.
  • Cross check the measured value of a quantity for improved accuracy. 
  • Use the instrument that has the highest precision.
  • It is suggested to pilot test measuring instruments for better accuracy.
  • Use multiple measures for the same construct. 
  • Note the measurements under controlled conditions.

Frequently Asked Questions – FAQs

What is meant by measurement error.

The difference between a measured quantity and its true value gives measurement error.

What are the types of errors?

The following are the types of errors:

The error seen due to the effect of the external conditions on the measurement is known as?

It is known as the environmental error.

Define absolute error?

A length was calculated to be 10.1 feet, but the absolute length was 10.5 feet. calculate the absolute error..

We know that, Absolute error = |VA-VE| Absolute error = 10.5-10.1 = 0.4 feet

Know in detail about absolute and relative error measurement . Visit BYJU’S – The Learning App and fall in love with learning!! 

types of errors in a science experiment

Put your understanding of this concept to test by answering a few MCQs. Click ‘Start Quiz’ to begin!

Select the correct answer and click on the “Finish” button Check your score and answers at the end of the quiz

Visit BYJU’S for all Physics related queries and study materials

Your result is as below

Request OTP on Voice Call

PHYSICS Related Links

Leave a Comment Cancel reply

Your Mobile number and Email id will not be published. Required fields are marked *

Post My Comment

types of errors in a science experiment

Register with BYJU'S & Download Free PDFs

Register with byju's & watch live videos.

Featured Topics

Featured series.

A series of random questions answered by Harvard experts.

Explore the Gazette

Read the latest.

Woman being served food from the Mediterranean diet.

Women who follow Mediterranean diet live longer

Statin Tablet in Close Up.

Harvard-led study IDs statin that may block pathway to some cancers

This illustration shows neurons with amyloid plaques, a hallmark of Alzheimer’s disease, in yellow.

New Alzheimer’s study suggests genetic cause of specific form of disease

Doctor monitoring medical equipment in the ICU.

High rate of diagnostic error found in ICU

Nationwide study pinpoints testing mistakes as most common cause.

A new study from Harvard-affiliated Brigham and Women’s Hospital, in collaboration with researchers at the University of California San Francisco, has shed light on the rate and impact of diagnostic errors in hospital settings.

In an analysis of electronic health records from 29 hospitals across the country of 2,428 patients who had either been transferred to an intensive care unit (ICU) or died in the hospital, the researchers found 550 patients (23 percent) experienced a diagnostic error, the majority of which were harmful to the patient. The researchers also determined the most common causes of diagnostic errors.

The study was published Monday in the journal JAMA Internal Medicine.

“We know diagnostic errors are dangerous and hospitals are obviously interested in reducing their frequency, but it’s much harder to do this when we don’t know what’s causing these errors or what their direct impact is on individual patients,” said senior author  Jeffrey L. Schnipper of the Brigham’s  Division of General Internal Medicine and Primary Care . “We found that diagnostic errors can largely be attributed to either errors in testing, or errors in assessing patients, and this knowledge gives us new opportunities to solve these problems.”

“It appears to be that only a minority of deaths in hospitals are linked to diagnostic errors, but even a single patient death that might have been prevented with a better diagnostic process is one death too many.”  Jeffrey L. Schnipper, Women and Brigham’s Hospital

Diagnostic errors are defined in medicine as any failure to either accurately explain a patient’s health problem or a failure to communicate that information to the patient. Some national efforts are currently underway to detect and address their causes, including  DECODE , a Diagnostic Centers of Excellence program at the Brigham that focuses on decreasing diagnostic errors in medical imaging by implementing and evaluating a highly resilient system for care planning and coordination, as well as a peer learning system for clinical providers. Other projects underway that involve BWH researchers will  address cases and causes of delayed diagnosis of cancer ,  explore how electronic health records contribute to diagnostic errors , and more.

To date, few studies have quantified the prevalence of diagnostic errors in hospitals or their most common underlying causes. In this study, cases were assessed for diagnostic errors by teams of two physicians who received extensive training in error adjudication and utilized multiple quality control steps. They found that 550 of the patients in their cohort (23 percent) experienced a diagnostic error in the hospital. Of these, 486 (17 percent of all patients) experienced some form of harm because of these errors. Of the 1,863 patients who died, the researchers judged that a diagnostic error was a contributing factor in 121 cases (6.6 percent).

“It appears to be that only a minority of deaths in hospitals are linked to diagnostic errors, but even a single patient death that might have been prevented with a better diagnostic process is one death too many,” said Schnipper.

The researchers found that most errors were attributable to errors in assessing patients, or errors in ordering and interpreting diagnostic tests.

“These two parts of the diagnostic process feed directly into each other,” said Schnipper. “If you don’t think of the correct possible diagnosis during your assessment of a patient, you’re not going to order the right tests. And if you order the wrong test or order the right test but misinterpret the result, this will inevitably change how you then assess a patient.”

While the research demonstrates the dangers that diagnostic errors can pose to patients, the researchers maintain that the rate of diagnostic errors in their specific population of patients, who all had experienced bad outcomes, does not represent the general rate across hospitals. The researchers are next exploring how health systems can implement surveillance systems to catch diagnostic errors as they occur, compare results across hospitals, and start  pilot testing  possible solutions.

“Our study does not tell us the overall frequency of diagnostic errors in the hospital, but it does tell us that there’s more we can be doing to prevent these types of errors from occurring,” said Schnipper.

Authors of the study include Andrew D. Auerbach, Tiffany M. Lee, Colin C. Hubbard, Sumant R. Ranji, Katie Raffel, Gilmer Valdes, John Boscardin, Anuj K. Dalal (BWH), Alyssa Harris, and Ellen Flynn for the UPSIDE Research Group. The authors declared no competing interests.

This study was supported by the Agency for Healthcare Research and Quality (R01HS027369).

Share this article

You might like.

Large study shows benefits against cancer, cardiovascular mortality, also identifies likely biological drivers of better health

Statin Tablet in Close Up.

Cholesterol-lowering drug suppresses chronic inflammation that creates dangerous cascade

This illustration shows neurons with amyloid plaques, a hallmark of Alzheimer’s disease, in yellow.

Findings eventually could pave way to earlier diagnosis, treatment, and affect search for new therapies

When should Harvard speak out?

Institutional Voice Working Group provides a roadmap in new report

Had a bad experience meditating? You're not alone.

Altered states of consciousness through yoga, mindfulness more common than thought and mostly beneficial, study finds — though clinicians ill-equipped to help those who struggle

Finding right mix on campus speech policies

Legal, political scholars discuss balancing personal safety, constitutional rights, academic freedom amid roiling protests, cultural shifts

IMAGES

  1. Types of Error

    types of errors in a science experiment

  2. PPT

    types of errors in a science experiment

  3. Types of Error

    types of errors in a science experiment

  4. PPT

    types of errors in a science experiment

  5. PPT

    types of errors in a science experiment

  6. Different Types of Errors in Physics and Their Examples

    types of errors in a science experiment

COMMENTS

  1. Sources of Error in Science Experiments

    Learn why all science experiments have error, how to calculate it, and the sources and types of errors you should report.

  2. Random vs. Systematic Error

    Random and systematic errors are types of measurement error, a difference between the observed and true values of something. FAQ About us . Our editors; Apply as editor; Team; Jobs; Contact ... In a controlled experiment, all variables other than the independent variable are held constant. 41. Extraneous Variables | Examples, Types & Controls ...

  3. PDF Introduction to Error and Uncertainty

    possible errors is an important issue in any experimental science. The conclusions we draw from the data, and especially the strength of those conclusions, will depend on

  4. 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 ...

  5. 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.

  6. Types of Error

    Random Errors: 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 of the experiment (such as ...

  7. PDF Measurement and Error Analysis

    Importance of Experimental Errors The scientific approach to understanding the world is based on a number of fundamental assumptions and techniques. Two of the most important are: • Experiments are reproducible. If you say experiment Xproduces result Y, I should be able to do experiment Xmyself and expect to get result Y.

  8. Types of Errors: Detection and Minimization

    Errors can be classified by source, magnitude, and sign. There are three types of errors: systematic, random, and gross. Systematic or determinate errors emerge from known sources and are reproducible during replicate measurements. Defective equipment and experiment design flaws are familiar sources of these errors.

  9. PDF ERROR ANALYSIS (UNCERTAINTY ANALYSIS)

    or. dy − dx. - These errors are much smaller. • In general if different errors are not correlated, are independent, the way to combine them is. dz =. dx2 + dy2. • This is true for random and bias errors. THE CASE OF Z = X - Y. • Suppose Z = X - Y is a number much smaller than X or Y.

  10. Systematic vs Random Error

    An old, stretched cloth measuring tape gives consistent, but different measurements than a new tape. Proportional errors of this type are called scale factor errors. Drift occurs when successive measurements become consistently higher or lower as time progresses. Electronic equipment is susceptible to drift.

  11. Experimental Errors and Error Analysis

    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: The word "precision" will

  12. PDF Experimental Uncertainties (Errors)

    experimental errors. Do not list all possible sources of errors there. Your goal is to identify only those significant for that experiment! For example, if the lab table is not perfectly leveled, then for the collision experiments (M6 - Impulse and Momentum) when the track is supposed to be horizontal, results will have a large, significant ...

  13. Issues with data and analyses: Errors, underlying themes, and ...

    These include collecting, analyzing, and reporting data. In each of these aspects, errors can and do occur. In this work, we first discuss the importance of focusing on statistical and data errors to continually improve the practice of science. We then describe underlying themes of the types of errors and postulate contributing factors.

  14. Bias and Sources of Error

    These types of errors are also known as "blunders" or "miscalculations" and they happen to everyone. In experimental inquires, these types of errors can occur due to: incorrect reading of instructions (e.g., 50 mL vs. 500 mL, sugar vs. salt, etc.); incorrect measuring (e.g., inches instead of cm, °F instead of °C, voltage instead of ...

  15. PDF Using This Checklist

    Lab ReportTypes of Experimental Errors. Using This Checklist Use this checklist as a preliminary guideline when thinking about and analyzing potential errors in your experiment. The list is a guide but is not comprehensive, so make sure that you check with your instructor about the different types of errors to pay attention to in your lab.

  16. Evaluating an experiment

    Systematic errors are due to problems with the equipment you used. For example, the balances you used may have been out by 0.1 g for every measurement. For example, the balances you used may have ...

  17. Errors in Measurement: Gross Errors, Systematic Errors and Random Errors

    Types of Errors. There are three types of errors that are classified based on the source they arise from; They are: Gross Errors; Random Errors; Systematic Errors; Gross Errors. This category basically takes into account human oversight and other mistakes while reading, recording, and readings.

  18. Obtaining, analysing and evaluating results

    Learn about valuable skills for doing an experiment, like creating hypotheses, identifying risks, and measuring and recording data accurately.

  19. Research assesses rates, causes of diagnostic errors

    In an analysis of electronic health records researchers found errors in assessing patients, or errors in ordering and interpreting ... Other projects underway that involve BWH ... but it does tell us that there's more we can be doing to prevent these types of errors from occurring," said Schnipper. Authors of the study include Andrew D ...