Select a year to see courses

Learn online or on-campus during the term or school holidays

  • Maths Acceleration
  • English Advanced
  • Maths Standard
  • Maths Advanced
  • Maths Extension 1
  • English Standard
  • Maths Extension 2

Get HSC exam ready in just a week

  • UCAT Exam Preparation

Select a year to see available courses

  • English Units 1/2
  • Maths Methods Units 1/2
  • Biology Units 1/2
  • Chemistry Units 1/2
  • Physics Units 1/2
  • English Units 3/4
  • Maths Methods Units 3/4
  • Biology Unit 3/4
  • Chemistry Unit 3/4
  • Physics Unit 3/4
  • VIC UCAT Preparation Course
  • Matrix Learning Methods
  • Matrix+ Online Courses
  • Matrix Term Courses
  • Matrix Holiday Courses
  • Campus overview
  • Castle Hill
  • Strathfield
  • Sydney City
  • Liverpool (Opening soon)
  • Year 3 NAPLAN Guide
  • OC Test Guide
  • Selective Schools Guide
  • NSW Primary School Rankings
  • NSW High School Rankings
  • NSW High Schools Guide
  • VIC School Rankings
  • ATAR & Scaling Guide
  • HSC Study Planning Kit
  • Student Success Secrets
  • Reading List
  • Year 6 English
  • Year 7 & 8 English
  • Year 9 English
  • Year 10 English
  • Year 11 English Standard
  • Year 11 English Advanced
  • Year 12 English Standard
  • Year 12 English Advanced
  • HSC English Skills
  • How To Write An Essay
  • How to Analyse Poetry
  • English Techniques Toolkit
  • Year 7 Maths
  • Year 8 Maths
  • Year 9 Maths
  • Year 10 Maths
  • Year 11 Maths Advanced
  • Year 11 Maths Extension 1
  • Year 12 Maths Standard 2
  • Year 12 Maths Advanced
  • Year 12 Maths Extension 1
  • Year 12 Maths Extension 2

Science guides to help you get ahead

  • Year 11 Biology
  • Year 11 Chemistry
  • Year 11 Physics
  • Year 12 Biology
  • Year 12 Chemistry
  • Year 12 Physics
  • Physics Practical Skills
  • Periodic Table
  • Set Location
  • 1300 008 008
  • 1300 634 117

Welcome to Matrix Education

To ensure we are showing you the most relevant content, please select your location below.

How to Write a Scientific Report | Step-by-Step Guide

Got to document an experiment but don't know how? In this post, we'll guide you step-by-step through how to write a scientific report and provide you with an example.

' src=

Get free study tips and resources delivered to your inbox.

Join 75,893 students who already have a head start.

" * " indicates required fields

You might also like

  • How To Do Essay Scaffolding Drills And Boost Your Essay Marks
  • How to Write a T.E.E. Statement
  • The Top 5 Persuasive Techniques for Speeches
  • How Parents Can Limit Study Distractions
  • Overcoming an English Slump in Years 9 and 10

Related courses

Year 9 science, year 10 science.

Is your teacher expecting you to write an experimental report for every class experiment? Are you still unsure about how to write a scientific report properly? Don’t fear! We’ll guide you through all parts of a scientific report, step-by-step.

How to write a scientific report:

  • What is a scientific report
  • General rules for writing Scientific reports
  • Introduction/Background information
  • Risk assessment

What is a scientific report?

A scientific report documents all aspects of an experimental investigation. This includes:

  • The aim of the experiment
  • The hypothesis
  • An introduction to the relevant background theory
  • The methods used
  • The results
  • A discussion of the results
  • The conclusion

Scientific reports allow their readers to understand the experiment without doing it themselves. These reports also give others the opportunity to check the methodology of the experiment to ensure the validity of the results.

scientific report essay example

Download your free experimental report template

A template to teach you to write perfect science reports

Done! Your download has been emailed.

Please allow a few minutes for it to land in your inbox.

scientific report essay example

We take your privacy seriously. T&Cs  and  Privacy Policy .

A scientific report is written in several stages. We write the introduction, aim, and hypothesis before performing the experiment, record the results during the experiment, and complete the discussion and conclusions after the experiment.

But, before we delve deeper into how to write a scientific report, we need to have a science experiment to write about! Choose an experiment to use as an example from our 7 Simple Experiments You Can Do At Home article.

blog-how-to-write-a-scientific-report-experiment

General rules for writing scientific reports

Learning how to write a scientific report is different from writing English essays or speeches.

You have to use:

  • Passive voice (which you should avoid when writing for other subjects like English!)
  • Past-tense language
  • Headings and subheadings
  • A pencil to draw scientific diagrams and graphs
  • Simple and clear lines for scientific diagrams
  • Tables and graphs where necessary

Structure of scientific reports

Now that you know the general rules on how to write scientific reports, let’s look at the conventions for their structure!

The title should simply introduce what your experiment is about.

The Role of Light in Photosynthesis

2. Introduction/Background information

Write a paragraph that gives your readers background information to understand your experiment.

This includes explaining scientific theories, processes, and other related knowledge.

Photosynthesis is a vital process for life. It occurs when plants intake carbon dioxide, water, and light, and results in the production of glucose and water. The light required for photosynthesis is absorbed by chlorophyll, the green pigment of plants, which is contained in the chloroplasts.

The glucose produced through photosynthesis is stored as starch, which is used as an energy source for the plant and its consumers.

The presence of starch in the leaves of a plant indicates that photosynthesis has occurred.

blog-how-to-write-a-scientific-report-photosynthesis

The aim tells us what is going to be tested in the experiment. This should be short, concise, and clear.

The aim of the experiment is to test whether light is required for photosynthesis to occur.

4. Hypothesis

The hypothesis is what you predict the outcome of the experiment will be. You have to use background information to make an educated prediction.

It is predicted that photosynthesis will occur only in leaves that are exposed to light and not in leaves that are not exposed to light. This will be indicated by the presence or absence of starch in the leaves.

5. Risk assessment

Identify the hazards related to the experiment and explain how you would prevent or minimise the risks. A hazard is something that can cause harm, and a risk is how likely it is that harm will occur because of the hazard.

A table is an excellent way to present your risk assessment.

Remember, you have to specify the  type of harm that can occur because of the hazard. It’s not enough to simply identify the hazard.

  • Do not write:  “Scissors are sharp.”
  • Instead, write: “Scissors are sharp and can cause injury.”
Scissors are sharp and can cause injuryLowUse the scissors correctly and store them after use.

Wear closed, durable shoes to prevent injury from falling sharp instruments.

Methylated spirits are highly flammable and can cause burns or fires.LowBefore using methylated spirits, ensure that all ignition sources such as Bunsen burners and matches are extinguished. Minimise the volume of methylated spirits used.

blog-how-to-write-a-scientific-report-photosynthesis-risk

The method has 3 parts:

  • A list of every material used
  • Steps of what you did in the experiment
  • A scientific diagram of the experimental apparatus

Let’s break down what you need to do for each section.

6a. Materials

This must list every piece of equipment and material used in the experiment.

Remember, you need to specify the amount of each material you used.

  • 1 geranium plant
  • Aluminium foil
  • 2 test tubes
  • 1 test tube rack
  • 1 pair of scissors
  • 1 250 mL beaker
  • 1 pair of forceps
  • 1 10 mL measuring cylinder
  • Iodine solution (5 mL)
  • Methylated spirit (50 mL)
  • Boiling water
  • 2 Petri dishes

blog-how-to-write-a-scientific-report-photosynthesis-material

As a rule of thumb, you should write the method clearly and in a way that readers can repeat the experiment and get similar results.

Using a numbered list for the steps of your experimental procedure is much clearer than writing a whole paragraph of text.  The steps should:

  • Be written in a sequential order, based on when they were performed
  • Specify any equipment that was used
  • Specify the quantity of any materials that were used

You also need to use past tense and passive voice when you’re writing your method. Scientific reports are supposed to show the readers what you did in the experiment, not what you will do.

  • Aluminium foil was used to fully cover a leaf of the geranium plant. The plant was left in the sun for three days.
  • On the third day, the covered leaf and 1 non-covered leaf were collected from the plant. The foil was removed from the covered leaf, and a 1 cm square was cut from each leaf using a pair of scissors.
  • 150 mL of water was boiled in a kettle and poured into a 250 mL beaker.
  • Using forceps, the 1 cm square of covered leaf was placed into the beaker of boiling water for 2 minutes. It was then placed in a test tube labelled “dark”.
  • The water in the beaker was discarded and replaced with 150 mL of freshly boiled water.
  • Using forceps, the 1 cm square non-covered leaf was placed into the beaker of boiling water for 2 minutes. It was then placed in a test tube labelled “light”.
  • 5 mL of methylated spirit was measured with a measuring cylinder and poured into each test tube so that the leaves were fully covered.
  • The water in the beaker was replaced with 150 mL of freshly boiled water and both the “light” and “dark” test tubes were immersed in the beaker of boiling water for 5 minutes.
  • The leaves were collected from each test tube with forceps, rinsed under cold running water, and placed onto separate labelled Petri dishes.
  • 3 drops of iodine solution were added to each leaf.
  • Both Petri dishes were placed side by side and observations were recorded.
  • The experiment was repeated 5 times, and results were compared between different groups.

6c. Diagram

After you complete your steps, it’s time to draw your scientific diagrams! Here are some rules for drawing scientific diagrams:

  • Always use a pencil to draw your scientific diagrams.
  • Use simple, sharp, 2D lines and shapes to draw your diagram. Don’t draw 3D shapes or use shading.
  • Label everything in your diagram.
  • Use thin, straight lines to label your diagram. Don’t use arrows.
  • Ensure that the label lines touch the outline of the equipment you are labelling and do not cross over it or stop short of it.
  • The label lines should never cross over each other.
  • Use a ruler for any straight lines in your diagram.
  • Draw a sufficiently large diagram so all components can be seen clearly.

blog-how-to-write-a-scientific-report-scientific-diagram-photosynthesis

This is where you document the results of your experiment. The data that you record for your experiment will generally be qualitative and/or quantitative.

Qualitative data is data that relates to qualities and is based on observations (qualitative – quality). This type of data is descriptive and is recorded in words. For example, the colour changed from green to orange, or the liquid became hot.

Quantitative data refers to n umerical data (quantitative – quantity). This type of data is recorded using numbers and is either measured or counted. For example, the plant grew 5.2 cm, or there were 5 frogs.

The way you record your results is important. Most of the time, a table is the best way to do this.

Here are some rules to using tables

  • Use a pencil and a ruler to draw your table.
  • Draw neat and straight lines.
  • Ensure that the table is closed (connect all your lines).
  • Don’t cross your lines (erase any lines that stick out of the table).
  • Use appropriate columns and rows.
  • Properly name each column and row (including the units of measurement in brackets).
  • Do not write your units in the body of your table (units belong in the header).
  • Always include a title.

Note : If your results require calculations, clearly write each step.

Observations of the effects of light on the amount of starch in plant leaves.

Dark blue, purple and blackYes
Light-yellowNo

blog-how-to-write-a-scientific-report-photosynthesis-results

If quantitative data was recorded, the data is often also plotted on a graph.

8. Discussion

The discussion is where you analyse and interpret your results, and identify any experimental errors or possible areas of improvements.

You should divide your discussion as follows.

1. Trend in the results

Describe the ‘trend’ in your results. That is, the relationship you observed between your independent and dependent variables.

The independent variable is the variable that you’re changing in the experiment. In this experiment, it’s the amount of light that the leaves are exposed to.

The dependent variable is the variable that you’re measuring in the experiment. In this experiment, it’s the presence of starch in the leaves.

The presence of starch is indicated when the addition of iodine causes the leaf to turn dark purple. The results show that starch was present in the leaves that were exposed to light, while the leaves that were not exposed to light did not contain starch.

2. Scientific explanation

Give an explanation of the results using scientific knowledge, theories, and any other scientific resources you find.

As starch is produced during photosynthesis, these results show that light plays a key role in photosynthesis.

3. Validity 

Validity refers to whether or not your results are valid. This can be done by examining your variables.

VA lidity =  VA riables

Identify the independent, dependent, controlled variables, and the control experiment (if you have one).

The controlled variables are the variables that you keep the same across all tests e.g. the size of the leaf sample.

The control experiment is where you don’t apply an independent variable. It is untouched for the whole experiment.

Ensure that you never change more than one variable at a time!

The independent variable of the experiment was the amount of light that the leaves were exposed to (the covered and uncovered geranium leaf), while the dependent variable was the presence of starch. The controlled variables were the size of the leaf sample, the duration of the experiment, the amount of time the solutions were heated, and the amount of iodine solution used.

4. Reliability 

Identify how you ensured the reliability of the results.

RE liability = RE petition

Show that you repeated your experiments, cross-checked your results with other groups, or collated your results with the class.

The reliability of the results was ensured by repeating the experiment 5 times and comparing results with other groups. Since other groups obtained comparable results, the results are reliable.

5. Accuracy

You should discuss accuracy if your results are in the form of quantitative data (numerical), and there is an accepted value for the result.

Accuracy wouldn’t be discussed for our photosynthesis experiment as qualitative data was collected. But, it would be discussed if we were measuring gravity using a pendulum:

The measured value of gravity was 9.8 m/s 2 , which is in agreement with the accepted value of 9.8 m/s 2 .

6. Possible improvements 

Point out any errors or risks you found in the experiment and suggest a method to improve it.

If there are none, then suggest new ways to improve the experimental design, and/or minimise error and risks.

blog-how-to-write-a-scientific-report-improve

Possible improvements could be made by including control experiments. For example, testing whether the iodine solution turns dark purple when added to water or methylated spirits. This would help to ensure that the purple colour observed in the experiments is due to the presence of starch in the leaves rather than impurities.

9. Conclusion

State whether the aim was achieved and if your hypothesis was supported.

The aim of the investigation was achieved, and it was found that light is required for photosynthesis to occur. This was evidenced by the presence of starch in leaves that had been exposed to light, and the absence of starch in leaves that had been unexposed. These results support the proposed hypothesis.

Learn everything you need to know about Science

Gain the confidence and skills to ace Science!

Learn one term ahead of school with the Subject Matter Experts!

Learn more ?

CTA Isabelle Teaching Physics

Written by Matrix Science Team

' src=

© Matrix Education and www.matrix.edu.au, 2023. Unauthorised use and/or duplication of this material without express and written permission from this site’s author and/or owner is strictly prohibited. Excerpts and links may be used, provided that full and clear credit is given to Matrix Education and www.matrix.edu.au with appropriate and specific direction to the original content.

Year 9 Science tutoring at Matrix is known for helping students build a strong foundation before studying Biology, Chemistry or Physics in senior school.

Learning methods available

Level 9 Science tutoring program

Year 10 Science tutoring at Matrix is known for helping students build a strong foundation before studying Biology, Chemistry or Physics in Year 11 and 12.

Level 10 Science tutoring program for Victorian students

Related articles

Matrix Study

Helping My Child Study Effectively

Parents often feel unsure of how to help their children study. In this post, we give you tips to get started helping your child succeed.

blog-english-5-must-dos-to-write-a-paragraph-how-to-write-a-paragraph-hero

5 Must Dos for Writing a Paragraph

Writing paragraphs is just like constructing buildings from blocks. You put together different elements to create a paragraph! So, let's find out what these 5 important elements are!

scientific report essay example

The Ultimate Year 6 Final Exams Study Guide

Need help getting started with your study for your Year 6 finals? Don't worry! This article will help you get ready to ace them!

The Writing Center • University of North Carolina at Chapel Hill

Scientific Reports

What this handout is about.

This handout provides a general guide to writing reports about scientific research you’ve performed. In addition to describing the conventional rules about the format and content of a lab report, we’ll also attempt to convey why these rules exist, so you’ll get a clearer, more dependable idea of how to approach this writing situation. Readers of this handout may also find our handout on writing in the sciences useful.

Background and pre-writing

Why do we write research reports.

You did an experiment or study for your science class, and now you have to write it up for your teacher to review. You feel that you understood the background sufficiently, designed and completed the study effectively, obtained useful data, and can use those data to draw conclusions about a scientific process or principle. But how exactly do you write all that? What is your teacher expecting to see?

To take some of the guesswork out of answering these questions, try to think beyond the classroom setting. In fact, you and your teacher are both part of a scientific community, and the people who participate in this community tend to share the same values. As long as you understand and respect these values, your writing will likely meet the expectations of your audience—including your teacher.

So why are you writing this research report? The practical answer is “Because the teacher assigned it,” but that’s classroom thinking. Generally speaking, people investigating some scientific hypothesis have a responsibility to the rest of the scientific world to report their findings, particularly if these findings add to or contradict previous ideas. The people reading such reports have two primary goals:

  • They want to gather the information presented.
  • They want to know that the findings are legitimate.

Your job as a writer, then, is to fulfill these two goals.

How do I do that?

Good question. Here is the basic format scientists have designed for research reports:

  • Introduction

Methods and Materials

This format, sometimes called “IMRAD,” may take slightly different shapes depending on the discipline or audience; some ask you to include an abstract or separate section for the hypothesis, or call the Discussion section “Conclusions,” or change the order of the sections (some professional and academic journals require the Methods section to appear last). Overall, however, the IMRAD format was devised to represent a textual version of the scientific method.

The scientific method, you’ll probably recall, involves developing a hypothesis, testing it, and deciding whether your findings support the hypothesis. In essence, the format for a research report in the sciences mirrors the scientific method but fleshes out the process a little. Below, you’ll find a table that shows how each written section fits into the scientific method and what additional information it offers the reader.

states your hypothesis explains how you derived that hypothesis and how it connects to previous research; gives the purpose of the experiment/study
details how you tested your hypothesis clarifies why you performed your study in that particular way
provides raw (i.e., uninterpreted) data collected (perhaps) expresses the data in table form, as an easy-to-read figure, or as percentages/ratios
considers whether the data you obtained support the hypothesis explores the implications of your finding and judges the potential limitations of your experimental design

Thinking of your research report as based on the scientific method, but elaborated in the ways described above, may help you to meet your audience’s expectations successfully. We’re going to proceed by explicitly connecting each section of the lab report to the scientific method, then explaining why and how you need to elaborate that section.

Although this handout takes each section in the order in which it should be presented in the final report, you may for practical reasons decide to compose sections in another order. For example, many writers find that composing their Methods and Results before the other sections helps to clarify their idea of the experiment or study as a whole. You might consider using each assignment to practice different approaches to drafting the report, to find the order that works best for you.

What should I do before drafting the lab report?

The best way to prepare to write the lab report is to make sure that you fully understand everything you need to about the experiment. Obviously, if you don’t quite know what went on during the lab, you’re going to find it difficult to explain the lab satisfactorily to someone else. To make sure you know enough to write the report, complete the following steps:

  • What are we going to do in this lab? (That is, what’s the procedure?)
  • Why are we going to do it that way?
  • What are we hoping to learn from this experiment?
  • Why would we benefit from this knowledge?
  • Consult your lab supervisor as you perform the lab. If you don’t know how to answer one of the questions above, for example, your lab supervisor will probably be able to explain it to you (or, at least, help you figure it out).
  • Plan the steps of the experiment carefully with your lab partners. The less you rush, the more likely it is that you’ll perform the experiment correctly and record your findings accurately. Also, take some time to think about the best way to organize the data before you have to start putting numbers down. If you can design a table to account for the data, that will tend to work much better than jotting results down hurriedly on a scrap piece of paper.
  • Record the data carefully so you get them right. You won’t be able to trust your conclusions if you have the wrong data, and your readers will know you messed up if the other three people in your group have “97 degrees” and you have “87.”
  • Consult with your lab partners about everything you do. Lab groups often make one of two mistakes: two people do all the work while two have a nice chat, or everybody works together until the group finishes gathering the raw data, then scrams outta there. Collaborate with your partners, even when the experiment is “over.” What trends did you observe? Was the hypothesis supported? Did you all get the same results? What kind of figure should you use to represent your findings? The whole group can work together to answer these questions.
  • Consider your audience. You may believe that audience is a non-issue: it’s your lab TA, right? Well, yes—but again, think beyond the classroom. If you write with only your lab instructor in mind, you may omit material that is crucial to a complete understanding of your experiment, because you assume the instructor knows all that stuff already. As a result, you may receive a lower grade, since your TA won’t be sure that you understand all the principles at work. Try to write towards a student in the same course but a different lab section. That student will have a fair degree of scientific expertise but won’t know much about your experiment particularly. Alternatively, you could envision yourself five years from now, after the reading and lectures for this course have faded a bit. What would you remember, and what would you need explained more clearly (as a refresher)?

Once you’ve completed these steps as you perform the experiment, you’ll be in a good position to draft an effective lab report.

Introductions

How do i write a strong introduction.

For the purposes of this handout, we’ll consider the Introduction to contain four basic elements: the purpose, the scientific literature relevant to the subject, the hypothesis, and the reasons you believed your hypothesis viable. Let’s start by going through each element of the Introduction to clarify what it covers and why it’s important. Then we can formulate a logical organizational strategy for the section.

The inclusion of the purpose (sometimes called the objective) of the experiment often confuses writers. The biggest misconception is that the purpose is the same as the hypothesis. Not quite. We’ll get to hypotheses in a minute, but basically they provide some indication of what you expect the experiment to show. The purpose is broader, and deals more with what you expect to gain through the experiment. In a professional setting, the hypothesis might have something to do with how cells react to a certain kind of genetic manipulation, but the purpose of the experiment is to learn more about potential cancer treatments. Undergraduate reports don’t often have this wide-ranging a goal, but you should still try to maintain the distinction between your hypothesis and your purpose. In a solubility experiment, for example, your hypothesis might talk about the relationship between temperature and the rate of solubility, but the purpose is probably to learn more about some specific scientific principle underlying the process of solubility.

For starters, most people say that you should write out your working hypothesis before you perform the experiment or study. Many beginning science students neglect to do so and find themselves struggling to remember precisely which variables were involved in the process or in what way the researchers felt that they were related. Write your hypothesis down as you develop it—you’ll be glad you did.

As for the form a hypothesis should take, it’s best not to be too fancy or complicated; an inventive style isn’t nearly so important as clarity here. There’s nothing wrong with beginning your hypothesis with the phrase, “It was hypothesized that . . .” Be as specific as you can about the relationship between the different objects of your study. In other words, explain that when term A changes, term B changes in this particular way. Readers of scientific writing are rarely content with the idea that a relationship between two terms exists—they want to know what that relationship entails.

Not a hypothesis:

“It was hypothesized that there is a significant relationship between the temperature of a solvent and the rate at which a solute dissolves.”

Hypothesis:

“It was hypothesized that as the temperature of a solvent increases, the rate at which a solute will dissolve in that solvent increases.”

Put more technically, most hypotheses contain both an independent and a dependent variable. The independent variable is what you manipulate to test the reaction; the dependent variable is what changes as a result of your manipulation. In the example above, the independent variable is the temperature of the solvent, and the dependent variable is the rate of solubility. Be sure that your hypothesis includes both variables.

Justify your hypothesis

You need to do more than tell your readers what your hypothesis is; you also need to assure them that this hypothesis was reasonable, given the circumstances. In other words, use the Introduction to explain that you didn’t just pluck your hypothesis out of thin air. (If you did pluck it out of thin air, your problems with your report will probably extend beyond using the appropriate format.) If you posit that a particular relationship exists between the independent and the dependent variable, what led you to believe your “guess” might be supported by evidence?

Scientists often refer to this type of justification as “motivating” the hypothesis, in the sense that something propelled them to make that prediction. Often, motivation includes what we already know—or rather, what scientists generally accept as true (see “Background/previous research” below). But you can also motivate your hypothesis by relying on logic or on your own observations. If you’re trying to decide which solutes will dissolve more rapidly in a solvent at increased temperatures, you might remember that some solids are meant to dissolve in hot water (e.g., bouillon cubes) and some are used for a function precisely because they withstand higher temperatures (they make saucepans out of something). Or you can think about whether you’ve noticed sugar dissolving more rapidly in your glass of iced tea or in your cup of coffee. Even such basic, outside-the-lab observations can help you justify your hypothesis as reasonable.

Background/previous research

This part of the Introduction demonstrates to the reader your awareness of how you’re building on other scientists’ work. If you think of the scientific community as engaging in a series of conversations about various topics, then you’ll recognize that the relevant background material will alert the reader to which conversation you want to enter.

Generally speaking, authors writing journal articles use the background for slightly different purposes than do students completing assignments. Because readers of academic journals tend to be professionals in the field, authors explain the background in order to permit readers to evaluate the study’s pertinence for their own work. You, on the other hand, write toward a much narrower audience—your peers in the course or your lab instructor—and so you must demonstrate that you understand the context for the (presumably assigned) experiment or study you’ve completed. For example, if your professor has been talking about polarity during lectures, and you’re doing a solubility experiment, you might try to connect the polarity of a solid to its relative solubility in certain solvents. In any event, both professional researchers and undergraduates need to connect the background material overtly to their own work.

Organization of this section

Most of the time, writers begin by stating the purpose or objectives of their own work, which establishes for the reader’s benefit the “nature and scope of the problem investigated” (Day 1994). Once you have expressed your purpose, you should then find it easier to move from the general purpose, to relevant material on the subject, to your hypothesis. In abbreviated form, an Introduction section might look like this:

“The purpose of the experiment was to test conventional ideas about solubility in the laboratory [purpose] . . . According to Whitecoat and Labrat (1999), at higher temperatures the molecules of solvents move more quickly . . . We know from the class lecture that molecules moving at higher rates of speed collide with one another more often and thus break down more easily [background material/motivation] . . . Thus, it was hypothesized that as the temperature of a solvent increases, the rate at which a solute will dissolve in that solvent increases [hypothesis].”

Again—these are guidelines, not commandments. Some writers and readers prefer different structures for the Introduction. The one above merely illustrates a common approach to organizing material.

How do I write a strong Materials and Methods section?

As with any piece of writing, your Methods section will succeed only if it fulfills its readers’ expectations, so you need to be clear in your own mind about the purpose of this section. Let’s review the purpose as we described it above: in this section, you want to describe in detail how you tested the hypothesis you developed and also to clarify the rationale for your procedure. In science, it’s not sufficient merely to design and carry out an experiment. Ultimately, others must be able to verify your findings, so your experiment must be reproducible, to the extent that other researchers can follow the same procedure and obtain the same (or similar) results.

Here’s a real-world example of the importance of reproducibility. In 1989, physicists Stanley Pons and Martin Fleischman announced that they had discovered “cold fusion,” a way of producing excess heat and power without the nuclear radiation that accompanies “hot fusion.” Such a discovery could have great ramifications for the industrial production of energy, so these findings created a great deal of interest. When other scientists tried to duplicate the experiment, however, they didn’t achieve the same results, and as a result many wrote off the conclusions as unjustified (or worse, a hoax). To this day, the viability of cold fusion is debated within the scientific community, even though an increasing number of researchers believe it possible. So when you write your Methods section, keep in mind that you need to describe your experiment well enough to allow others to replicate it exactly.

With these goals in mind, let’s consider how to write an effective Methods section in terms of content, structure, and style.

Sometimes the hardest thing about writing this section isn’t what you should talk about, but what you shouldn’t talk about. Writers often want to include the results of their experiment, because they measured and recorded the results during the course of the experiment. But such data should be reserved for the Results section. In the Methods section, you can write that you recorded the results, or how you recorded the results (e.g., in a table), but you shouldn’t write what the results were—not yet. Here, you’re merely stating exactly how you went about testing your hypothesis. As you draft your Methods section, ask yourself the following questions:

  • How much detail? Be precise in providing details, but stay relevant. Ask yourself, “Would it make any difference if this piece were a different size or made from a different material?” If not, you probably don’t need to get too specific. If so, you should give as many details as necessary to prevent this experiment from going awry if someone else tries to carry it out. Probably the most crucial detail is measurement; you should always quantify anything you can, such as time elapsed, temperature, mass, volume, etc.
  • Rationale: Be sure that as you’re relating your actions during the experiment, you explain your rationale for the protocol you developed. If you capped a test tube immediately after adding a solute to a solvent, why did you do that? (That’s really two questions: why did you cap it, and why did you cap it immediately?) In a professional setting, writers provide their rationale as a way to explain their thinking to potential critics. On one hand, of course, that’s your motivation for talking about protocol, too. On the other hand, since in practical terms you’re also writing to your teacher (who’s seeking to evaluate how well you comprehend the principles of the experiment), explaining the rationale indicates that you understand the reasons for conducting the experiment in that way, and that you’re not just following orders. Critical thinking is crucial—robots don’t make good scientists.
  • Control: Most experiments will include a control, which is a means of comparing experimental results. (Sometimes you’ll need to have more than one control, depending on the number of hypotheses you want to test.) The control is exactly the same as the other items you’re testing, except that you don’t manipulate the independent variable-the condition you’re altering to check the effect on the dependent variable. For example, if you’re testing solubility rates at increased temperatures, your control would be a solution that you didn’t heat at all; that way, you’ll see how quickly the solute dissolves “naturally” (i.e., without manipulation), and you’ll have a point of reference against which to compare the solutions you did heat.

Describe the control in the Methods section. Two things are especially important in writing about the control: identify the control as a control, and explain what you’re controlling for. Here is an example:

“As a control for the temperature change, we placed the same amount of solute in the same amount of solvent, and let the solution stand for five minutes without heating it.”

Structure and style

Organization is especially important in the Methods section of a lab report because readers must understand your experimental procedure completely. Many writers are surprised by the difficulty of conveying what they did during the experiment, since after all they’re only reporting an event, but it’s often tricky to present this information in a coherent way. There’s a fairly standard structure you can use to guide you, and following the conventions for style can help clarify your points.

  • Subsections: Occasionally, researchers use subsections to report their procedure when the following circumstances apply: 1) if they’ve used a great many materials; 2) if the procedure is unusually complicated; 3) if they’ve developed a procedure that won’t be familiar to many of their readers. Because these conditions rarely apply to the experiments you’ll perform in class, most undergraduate lab reports won’t require you to use subsections. In fact, many guides to writing lab reports suggest that you try to limit your Methods section to a single paragraph.
  • Narrative structure: Think of this section as telling a story about a group of people and the experiment they performed. Describe what you did in the order in which you did it. You may have heard the old joke centered on the line, “Disconnect the red wire, but only after disconnecting the green wire,” where the person reading the directions blows everything to kingdom come because the directions weren’t in order. We’re used to reading about events chronologically, and so your readers will generally understand what you did if you present that information in the same way. Also, since the Methods section does generally appear as a narrative (story), you want to avoid the “recipe” approach: “First, take a clean, dry 100 ml test tube from the rack. Next, add 50 ml of distilled water.” You should be reporting what did happen, not telling the reader how to perform the experiment: “50 ml of distilled water was poured into a clean, dry 100 ml test tube.” Hint: most of the time, the recipe approach comes from copying down the steps of the procedure from your lab manual, so you may want to draft the Methods section initially without consulting your manual. Later, of course, you can go back and fill in any part of the procedure you inadvertently overlooked.
  • Past tense: Remember that you’re describing what happened, so you should use past tense to refer to everything you did during the experiment. Writers are often tempted to use the imperative (“Add 5 g of the solid to the solution”) because that’s how their lab manuals are worded; less frequently, they use present tense (“5 g of the solid are added to the solution”). Instead, remember that you’re talking about an event which happened at a particular time in the past, and which has already ended by the time you start writing, so simple past tense will be appropriate in this section (“5 g of the solid were added to the solution” or “We added 5 g of the solid to the solution”).
  • Active: We heated the solution to 80°C. (The subject, “we,” performs the action, heating.)
  • Passive: The solution was heated to 80°C. (The subject, “solution,” doesn’t do the heating–it is acted upon, not acting.)

Increasingly, especially in the social sciences, using first person and active voice is acceptable in scientific reports. Most readers find that this style of writing conveys information more clearly and concisely. This rhetorical choice thus brings two scientific values into conflict: objectivity versus clarity. Since the scientific community hasn’t reached a consensus about which style it prefers, you may want to ask your lab instructor.

How do I write a strong Results section?

Here’s a paradox for you. The Results section is often both the shortest (yay!) and most important (uh-oh!) part of your report. Your Materials and Methods section shows how you obtained the results, and your Discussion section explores the significance of the results, so clearly the Results section forms the backbone of the lab report. This section provides the most critical information about your experiment: the data that allow you to discuss how your hypothesis was or wasn’t supported. But it doesn’t provide anything else, which explains why this section is generally shorter than the others.

Before you write this section, look at all the data you collected to figure out what relates significantly to your hypothesis. You’ll want to highlight this material in your Results section. Resist the urge to include every bit of data you collected, since perhaps not all are relevant. Also, don’t try to draw conclusions about the results—save them for the Discussion section. In this section, you’re reporting facts. Nothing your readers can dispute should appear in the Results section.

Most Results sections feature three distinct parts: text, tables, and figures. Let’s consider each part one at a time.

This should be a short paragraph, generally just a few lines, that describes the results you obtained from your experiment. In a relatively simple experiment, one that doesn’t produce a lot of data for you to repeat, the text can represent the entire Results section. Don’t feel that you need to include lots of extraneous detail to compensate for a short (but effective) text; your readers appreciate discrimination more than your ability to recite facts. In a more complex experiment, you may want to use tables and/or figures to help guide your readers toward the most important information you gathered. In that event, you’ll need to refer to each table or figure directly, where appropriate:

“Table 1 lists the rates of solubility for each substance”

“Solubility increased as the temperature of the solution increased (see Figure 1).”

If you do use tables or figures, make sure that you don’t present the same material in both the text and the tables/figures, since in essence you’ll just repeat yourself, probably annoying your readers with the redundancy of your statements.

Feel free to describe trends that emerge as you examine the data. Although identifying trends requires some judgment on your part and so may not feel like factual reporting, no one can deny that these trends do exist, and so they properly belong in the Results section. Example:

“Heating the solution increased the rate of solubility of polar solids by 45% but had no effect on the rate of solubility in solutions containing non-polar solids.”

This point isn’t debatable—you’re just pointing out what the data show.

As in the Materials and Methods section, you want to refer to your data in the past tense, because the events you recorded have already occurred and have finished occurring. In the example above, note the use of “increased” and “had,” rather than “increases” and “has.” (You don’t know from your experiment that heating always increases the solubility of polar solids, but it did that time.)

You shouldn’t put information in the table that also appears in the text. You also shouldn’t use a table to present irrelevant data, just to show you did collect these data during the experiment. Tables are good for some purposes and situations, but not others, so whether and how you’ll use tables depends upon what you need them to accomplish.

Tables are useful ways to show variation in data, but not to present a great deal of unchanging measurements. If you’re dealing with a scientific phenomenon that occurs only within a certain range of temperatures, for example, you don’t need to use a table to show that the phenomenon didn’t occur at any of the other temperatures. How useful is this table?

A table labeled Effect of Temperature on Rate of Solubility with temperature of solvent values in 10-degree increments from -20 degrees Celsius to 80 degrees Celsius that does not show a corresponding rate of solubility value until 50 degrees Celsius.

As you can probably see, no solubility was observed until the trial temperature reached 50°C, a fact that the text part of the Results section could easily convey. The table could then be limited to what happened at 50°C and higher, thus better illustrating the differences in solubility rates when solubility did occur.

As a rule, try not to use a table to describe any experimental event you can cover in one sentence of text. Here’s an example of an unnecessary table from How to Write and Publish a Scientific Paper , by Robert A. Day:

A table labeled Oxygen requirements of various species of Streptomyces showing the names of organisms and two columns that indicate growth under aerobic conditions and growth under anaerobic conditions with a plus or minus symbol for each organism in the growth columns to indicate value.

As Day notes, all the information in this table can be summarized in one sentence: “S. griseus, S. coelicolor, S. everycolor, and S. rainbowenski grew under aerobic conditions, whereas S. nocolor and S. greenicus required anaerobic conditions.” Most readers won’t find the table clearer than that one sentence.

When you do have reason to tabulate material, pay attention to the clarity and readability of the format you use. Here are a few tips:

  • Number your table. Then, when you refer to the table in the text, use that number to tell your readers which table they can review to clarify the material.
  • Give your table a title. This title should be descriptive enough to communicate the contents of the table, but not so long that it becomes difficult to follow. The titles in the sample tables above are acceptable.
  • Arrange your table so that readers read vertically, not horizontally. For the most part, this rule means that you should construct your table so that like elements read down, not across. Think about what you want your readers to compare, and put that information in the column (up and down) rather than in the row (across). Usually, the point of comparison will be the numerical data you collect, so especially make sure you have columns of numbers, not rows.Here’s an example of how drastically this decision affects the readability of your table (from A Short Guide to Writing about Chemistry , by Herbert Beall and John Trimbur). Look at this table, which presents the relevant data in horizontal rows:

A table labeled Boyle's Law Experiment: Measuring Volume as a Function of Pressure that presents the trial number, length of air sample in millimeters, and height difference in inches of mercury, each of which is presented in rows horizontally.

It’s a little tough to see the trends that the author presumably wants to present in this table. Compare this table, in which the data appear vertically:

A table labeled Boyle's Law Experiment: Measuring Volume as a Function of Pressure that presents the trial number, length of air sample in millimeters, and height difference in inches of mercury, each of which is presented in columns vertically.

The second table shows how putting like elements in a vertical column makes for easier reading. In this case, the like elements are the measurements of length and height, over five trials–not, as in the first table, the length and height measurements for each trial.

  • Make sure to include units of measurement in the tables. Readers might be able to guess that you measured something in millimeters, but don’t make them try.
1058
432
7
  • Don’t use vertical lines as part of the format for your table. This convention exists because journals prefer not to have to reproduce these lines because the tables then become more expensive to print. Even though it’s fairly unlikely that you’ll be sending your Biology 11 lab report to Science for publication, your readers still have this expectation. Consequently, if you use the table-drawing option in your word-processing software, choose the option that doesn’t rely on a “grid” format (which includes vertical lines).

How do I include figures in my report?

Although tables can be useful ways of showing trends in the results you obtained, figures (i.e., illustrations) can do an even better job of emphasizing such trends. Lab report writers often use graphic representations of the data they collected to provide their readers with a literal picture of how the experiment went.

When should you use a figure?

Remember the circumstances under which you don’t need a table: when you don’t have a great deal of data or when the data you have don’t vary a lot. Under the same conditions, you would probably forgo the figure as well, since the figure would be unlikely to provide your readers with an additional perspective. Scientists really don’t like their time wasted, so they tend not to respond favorably to redundancy.

If you’re trying to decide between using a table and creating a figure to present your material, consider the following a rule of thumb. The strength of a table lies in its ability to supply large amounts of exact data, whereas the strength of a figure is its dramatic illustration of important trends within the experiment. If you feel that your readers won’t get the full impact of the results you obtained just by looking at the numbers, then a figure might be appropriate.

Of course, an undergraduate class may expect you to create a figure for your lab experiment, if only to make sure that you can do so effectively. If this is the case, then don’t worry about whether to use figures or not—concentrate instead on how best to accomplish your task.

Figures can include maps, photographs, pen-and-ink drawings, flow charts, bar graphs, and section graphs (“pie charts”). But the most common figure by far, especially for undergraduates, is the line graph, so we’ll focus on that type in this handout.

At the undergraduate level, you can often draw and label your graphs by hand, provided that the result is clear, legible, and drawn to scale. Computer technology has, however, made creating line graphs a lot easier. Most word-processing software has a number of functions for transferring data into graph form; many scientists have found Microsoft Excel, for example, a helpful tool in graphing results. If you plan on pursuing a career in the sciences, it may be well worth your while to learn to use a similar program.

Computers can’t, however, decide for you how your graph really works; you have to know how to design your graph to meet your readers’ expectations. Here are some of these expectations:

  • Keep it as simple as possible. You may be tempted to signal the complexity of the information you gathered by trying to design a graph that accounts for that complexity. But remember the purpose of your graph: to dramatize your results in a manner that’s easy to see and grasp. Try not to make the reader stare at the graph for a half hour to find the important line among the mass of other lines. For maximum effectiveness, limit yourself to three to five lines per graph; if you have more data to demonstrate, use a set of graphs to account for it, rather than trying to cram it all into a single figure.
  • Plot the independent variable on the horizontal (x) axis and the dependent variable on the vertical (y) axis. Remember that the independent variable is the condition that you manipulated during the experiment and the dependent variable is the condition that you measured to see if it changed along with the independent variable. Placing the variables along their respective axes is mostly just a convention, but since your readers are accustomed to viewing graphs in this way, you’re better off not challenging the convention in your report.
  • Label each axis carefully, and be especially careful to include units of measure. You need to make sure that your readers understand perfectly well what your graph indicates.
  • Number and title your graphs. As with tables, the title of the graph should be informative but concise, and you should refer to your graph by number in the text (e.g., “Figure 1 shows the increase in the solubility rate as a function of temperature”).
  • Many editors of professional scientific journals prefer that writers distinguish the lines in their graphs by attaching a symbol to them, usually a geometric shape (triangle, square, etc.), and using that symbol throughout the curve of the line. Generally, readers have a hard time distinguishing dotted lines from dot-dash lines from straight lines, so you should consider staying away from this system. Editors don’t usually like different-colored lines within a graph because colors are difficult and expensive to reproduce; colors may, however, be great for your purposes, as long as you’re not planning to submit your paper to Nature. Use your discretion—try to employ whichever technique dramatizes the results most effectively.
  • Try to gather data at regular intervals, so the plot points on your graph aren’t too far apart. You can’t be sure of the arc you should draw between the plot points if the points are located at the far corners of the graph; over a fifteen-minute interval, perhaps the change occurred in the first or last thirty seconds of that period (in which case your straight-line connection between the points is misleading).
  • If you’re worried that you didn’t collect data at sufficiently regular intervals during your experiment, go ahead and connect the points with a straight line, but you may want to examine this problem as part of your Discussion section.
  • Make your graph large enough so that everything is legible and clearly demarcated, but not so large that it either overwhelms the rest of the Results section or provides a far greater range than you need to illustrate your point. If, for example, the seedlings of your plant grew only 15 mm during the trial, you don’t need to construct a graph that accounts for 100 mm of growth. The lines in your graph should more or less fill the space created by the axes; if you see that your data is confined to the lower left portion of the graph, you should probably re-adjust your scale.
  • If you create a set of graphs, make them the same size and format, including all the verbal and visual codes (captions, symbols, scale, etc.). You want to be as consistent as possible in your illustrations, so that your readers can easily make the comparisons you’re trying to get them to see.

How do I write a strong Discussion section?

The discussion section is probably the least formalized part of the report, in that you can’t really apply the same structure to every type of experiment. In simple terms, here you tell your readers what to make of the Results you obtained. If you have done the Results part well, your readers should already recognize the trends in the data and have a fairly clear idea of whether your hypothesis was supported. Because the Results can seem so self-explanatory, many students find it difficult to know what material to add in this last section.

Basically, the Discussion contains several parts, in no particular order, but roughly moving from specific (i.e., related to your experiment only) to general (how your findings fit in the larger scientific community). In this section, you will, as a rule, need to:

Explain whether the data support your hypothesis

  • Acknowledge any anomalous data or deviations from what you expected

Derive conclusions, based on your findings, about the process you’re studying

  • Relate your findings to earlier work in the same area (if you can)

Explore the theoretical and/or practical implications of your findings

Let’s look at some dos and don’ts for each of these objectives.

This statement is usually a good way to begin the Discussion, since you can’t effectively speak about the larger scientific value of your study until you’ve figured out the particulars of this experiment. You might begin this part of the Discussion by explicitly stating the relationships or correlations your data indicate between the independent and dependent variables. Then you can show more clearly why you believe your hypothesis was or was not supported. For example, if you tested solubility at various temperatures, you could start this section by noting that the rates of solubility increased as the temperature increased. If your initial hypothesis surmised that temperature change would not affect solubility, you would then say something like,

“The hypothesis that temperature change would not affect solubility was not supported by the data.”

Note: Students tend to view labs as practical tests of undeniable scientific truths. As a result, you may want to say that the hypothesis was “proved” or “disproved” or that it was “correct” or “incorrect.” These terms, however, reflect a degree of certainty that you as a scientist aren’t supposed to have. Remember, you’re testing a theory with a procedure that lasts only a few hours and relies on only a few trials, which severely compromises your ability to be sure about the “truth” you see. Words like “supported,” “indicated,” and “suggested” are more acceptable ways to evaluate your hypothesis.

Also, recognize that saying whether the data supported your hypothesis or not involves making a claim to be defended. As such, you need to show the readers that this claim is warranted by the evidence. Make sure that you’re very explicit about the relationship between the evidence and the conclusions you draw from it. This process is difficult for many writers because we don’t often justify conclusions in our regular lives. For example, you might nudge your friend at a party and whisper, “That guy’s drunk,” and once your friend lays eyes on the person in question, she might readily agree. In a scientific paper, by contrast, you would need to defend your claim more thoroughly by pointing to data such as slurred words, unsteady gait, and the lampshade-as-hat. In addition to pointing out these details, you would also need to show how (according to previous studies) these signs are consistent with inebriation, especially if they occur in conjunction with one another. To put it another way, tell your readers exactly how you got from point A (was the hypothesis supported?) to point B (yes/no).

Acknowledge any anomalous data, or deviations from what you expected

You need to take these exceptions and divergences into account, so that you qualify your conclusions sufficiently. For obvious reasons, your readers will doubt your authority if you (deliberately or inadvertently) overlook a key piece of data that doesn’t square with your perspective on what occurred. In a more philosophical sense, once you’ve ignored evidence that contradicts your claims, you’ve departed from the scientific method. The urge to “tidy up” the experiment is often strong, but if you give in to it you’re no longer performing good science.

Sometimes after you’ve performed a study or experiment, you realize that some part of the methods you used to test your hypothesis was flawed. In that case, it’s OK to suggest that if you had the chance to conduct your test again, you might change the design in this or that specific way in order to avoid such and such a problem. The key to making this approach work, though, is to be very precise about the weakness in your experiment, why and how you think that weakness might have affected your data, and how you would alter your protocol to eliminate—or limit the effects of—that weakness. Often, inexperienced researchers and writers feel the need to account for “wrong” data (remember, there’s no such animal), and so they speculate wildly about what might have screwed things up. These speculations include such factors as the unusually hot temperature in the room, or the possibility that their lab partners read the meters wrong, or the potentially defective equipment. These explanations are what scientists call “cop-outs,” or “lame”; don’t indicate that the experiment had a weakness unless you’re fairly certain that a) it really occurred and b) you can explain reasonably well how that weakness affected your results.

If, for example, your hypothesis dealt with the changes in solubility at different temperatures, then try to figure out what you can rationally say about the process of solubility more generally. If you’re doing an undergraduate lab, chances are that the lab will connect in some way to the material you’ve been covering either in lecture or in your reading, so you might choose to return to these resources as a way to help you think clearly about the process as a whole.

This part of the Discussion section is another place where you need to make sure that you’re not overreaching. Again, nothing you’ve found in one study would remotely allow you to claim that you now “know” something, or that something isn’t “true,” or that your experiment “confirmed” some principle or other. Hesitate before you go out on a limb—it’s dangerous! Use less absolutely conclusive language, including such words as “suggest,” “indicate,” “correspond,” “possibly,” “challenge,” etc.

Relate your findings to previous work in the field (if possible)

We’ve been talking about how to show that you belong in a particular community (such as biologists or anthropologists) by writing within conventions that they recognize and accept. Another is to try to identify a conversation going on among members of that community, and use your work to contribute to that conversation. In a larger philosophical sense, scientists can’t fully understand the value of their research unless they have some sense of the context that provoked and nourished it. That is, you have to recognize what’s new about your project (potentially, anyway) and how it benefits the wider body of scientific knowledge. On a more pragmatic level, especially for undergraduates, connecting your lab work to previous research will demonstrate to the TA that you see the big picture. You have an opportunity, in the Discussion section, to distinguish yourself from the students in your class who aren’t thinking beyond the barest facts of the study. Capitalize on this opportunity by putting your own work in context.

If you’re just beginning to work in the natural sciences (as a first-year biology or chemistry student, say), most likely the work you’ll be doing has already been performed and re-performed to a satisfactory degree. Hence, you could probably point to a similar experiment or study and compare/contrast your results and conclusions. More advanced work may deal with an issue that is somewhat less “resolved,” and so previous research may take the form of an ongoing debate, and you can use your own work to weigh in on that debate. If, for example, researchers are hotly disputing the value of herbal remedies for the common cold, and the results of your study suggest that Echinacea diminishes the symptoms but not the actual presence of the cold, then you might want to take some time in the Discussion section to recapitulate the specifics of the dispute as it relates to Echinacea as an herbal remedy. (Consider that you have probably already written in the Introduction about this debate as background research.)

This information is often the best way to end your Discussion (and, for all intents and purposes, the report). In argumentative writing generally, you want to use your closing words to convey the main point of your writing. This main point can be primarily theoretical (“Now that you understand this information, you’re in a better position to understand this larger issue”) or primarily practical (“You can use this information to take such and such an action”). In either case, the concluding statements help the reader to comprehend the significance of your project and your decision to write about it.

Since a lab report is argumentative—after all, you’re investigating a claim, and judging the legitimacy of that claim by generating and collecting evidence—it’s often a good idea to end your report with the same technique for establishing your main point. If you want to go the theoretical route, you might talk about the consequences your study has for the field or phenomenon you’re investigating. To return to the examples regarding solubility, you could end by reflecting on what your work on solubility as a function of temperature tells us (potentially) about solubility in general. (Some folks consider this type of exploration “pure” as opposed to “applied” science, although these labels can be problematic.) If you want to go the practical route, you could end by speculating about the medical, institutional, or commercial implications of your findings—in other words, answer the question, “What can this study help people to do?” In either case, you’re going to make your readers’ experience more satisfying, by helping them see why they spent their time learning what you had to teach them.

Works consulted

We consulted these works while writing this handout. This is not a comprehensive list of resources on the handout’s topic, and we encourage you to do your own research to find additional publications. Please do not use this list as a model for the format of your own reference list, as it may not match the citation style you are using. For guidance on formatting citations, please see the UNC Libraries citation tutorial . We revise these tips periodically and welcome feedback.

American Psychological Association. 2010. Publication Manual of the American Psychological Association . 6th ed. Washington, DC: American Psychological Association.

Beall, Herbert, and John Trimbur. 2001. A Short Guide to Writing About Chemistry , 2nd ed. New York: Longman.

Blum, Deborah, and Mary Knudson. 1997. A Field Guide for Science Writers: The Official Guide of the National Association of Science Writers . New York: Oxford University Press.

Booth, Wayne C., Gregory G. Colomb, Joseph M. Williams, Joseph Bizup, and William T. FitzGerald. 2016. The Craft of Research , 4th ed. Chicago: University of Chicago Press.

Briscoe, Mary Helen. 1996. Preparing Scientific Illustrations: A Guide to Better Posters, Presentations, and Publications , 2nd ed. New York: Springer-Verlag.

Council of Science Editors. 2014. Scientific Style and Format: The CSE Manual for Authors, Editors, and Publishers , 8th ed. Chicago & London: University of Chicago Press.

Davis, Martha. 2012. Scientific Papers and Presentations , 3rd ed. London: Academic Press.

Day, Robert A. 1994. How to Write and Publish a Scientific Paper , 4th ed. Phoenix: Oryx Press.

Porush, David. 1995. A Short Guide to Writing About Science . New York: Longman.

Williams, Joseph, and Joseph Bizup. 2017. Style: Lessons in Clarity and Grace , 12th ed. Boston: Pearson.

You may reproduce it for non-commercial use if you use the entire handout and attribute the source: The Writing Center, University of North Carolina at Chapel Hill

Make a Gift

American Psychological Association

Sample Papers

This page contains sample papers formatted in seventh edition APA Style. The sample papers show the format that authors should use to submit a manuscript for publication in a professional journal and that students should use to submit a paper to an instructor for a course assignment. You can download the Word files to use as templates and edit them as needed for the purposes of your own papers.

Most guidelines in the Publication Manual apply to both professional manuscripts and student papers. However, there are specific guidelines for professional papers versus student papers, including professional and student title page formats. All authors should check with the person or entity to whom they are submitting their paper (e.g., publisher or instructor) for guidelines that are different from or in addition to those specified by APA Style.

Sample papers from the Publication Manual

The following two sample papers were published in annotated form in the Publication Manual and are reproduced here as PDFs for your ease of use. The annotations draw attention to content and formatting and provide the relevant sections of the Publication Manual (7th ed.) to consult for more information.

  • Student sample paper with annotations (PDF, 5MB)
  • Professional sample paper with annotations (PDF, 2.7MB)

We also offer these sample papers in Microsoft Word (.docx) format with the annotations as comments to the text.

  • Student sample paper with annotations as comments (DOCX, 42KB)
  • Professional sample paper with annotations as comments (DOCX, 103KB)

Finally, we offer these sample papers in Microsoft Word (.docx) format without the annotations.

  • Student sample paper without annotations (DOCX, 36KB)
  • Professional sample paper without annotations (DOCX, 96KB)

Sample professional paper templates by paper type

These sample papers demonstrate APA Style formatting standards for different professional paper types. Professional papers can contain many different elements depending on the nature of the work. Authors seeking publication should refer to the journal’s instructions for authors or manuscript submission guidelines for specific requirements and/or sections to include.

  • Literature review professional paper template (DOCX, 47KB)
  • Mixed methods professional paper template (DOCX, 68KB)
  • Qualitative professional paper template (DOCX, 72KB)
  • Quantitative professional paper template (DOCX, 77KB)
  • Review professional paper template (DOCX, 112KB)

Sample papers are covered in the seventh edition APA Style manuals in the Publication Manual Chapter 2 and the Concise Guide Chapter 1

scientific report essay example

Related handouts

  • Heading Levels Template: Student Paper (PDF, 257KB)
  • Heading Levels Template: Professional Paper (PDF, 213KB)

Other instructional aids

  • Journal Article Reporting Standards (JARS)
  • APA Style Tutorials and Webinars
  • Handouts and Guides
  • Paper Format

View all instructional aids

Sample student paper templates by paper type

These sample papers demonstrate APA Style formatting standards for different student paper types. Students may write the same types of papers as professional authors (e.g., quantitative studies, literature reviews) or other types of papers for course assignments (e.g., reaction or response papers, discussion posts), dissertations, and theses.

APA does not set formal requirements for the nature or contents of an APA Style student paper. Students should follow the guidelines and requirements of their instructor, department, and/or institution when writing papers. For instance, an abstract and keywords are not required for APA Style student papers, although an instructor may request them in student papers that are longer or more complex. Specific questions about a paper being written for a course assignment should be directed to the instructor or institution assigning the paper.

  • Discussion post student paper template (DOCX, 31KB)
  • Literature review student paper template (DOCX, 37KB)
  • Quantitative study student paper template (DOCX, 53KB)

Sample papers in real life

Although published articles differ in format from manuscripts submitted for publication or student papers (e.g., different line spacing, font, margins, and column format), articles published in APA journals provide excellent demonstrations of APA Style in action.

APA journals began publishing papers in seventh edition APA Style in 2020. Professional authors should check the author submission guidelines for the journal to which they want to submit their paper for any journal-specific style requirements.

Credits for sample professional paper templates

Quantitative professional paper template: Adapted from “Fake News, Fast and Slow: Deliberation Reduces Belief in False (but Not True) News Headlines,” by B. Bago, D. G. Rand, and G. Pennycook, 2020, Journal of Experimental Psychology: General , 149 (8), pp. 1608–1613 ( https://doi.org/10.1037/xge0000729 ). Copyright 2020 by the American Psychological Association.

Qualitative professional paper template: Adapted from “‘My Smartphone Is an Extension of Myself’: A Holistic Qualitative Exploration of the Impact of Using a Smartphone,” by L. J. Harkin and D. Kuss, 2020, Psychology of Popular Media , 10 (1), pp. 28–38 ( https://doi.org/10.1037/ppm0000278 ). Copyright 2020 by the American Psychological Association.

Mixed methods professional paper template: Adapted from “‘I Am a Change Agent’: A Mixed Methods Analysis of Students’ Social Justice Value Orientation in an Undergraduate Community Psychology Course,” by D. X. Henderson, A. T. Majors, and M. Wright, 2019,  Scholarship of Teaching and Learning in Psychology , 7 (1), 68–80. ( https://doi.org/10.1037/stl0000171 ). Copyright 2019 by the American Psychological Association.

Literature review professional paper template: Adapted from “Rethinking Emotions in the Context of Infants’ Prosocial Behavior: The Role of Interest and Positive Emotions,” by S. I. Hammond and J. K. Drummond, 2019, Developmental Psychology , 55 (9), pp. 1882–1888 ( https://doi.org/10.1037/dev0000685 ). Copyright 2019 by the American Psychological Association.

Review professional paper template: Adapted from “Joining the Conversation: Teaching Students to Think and Communicate Like Scholars,” by E. L. Parks, 2022, Scholarship of Teaching and Learning in Psychology , 8 (1), pp. 70–78 ( https://doi.org/10.1037/stl0000193 ). Copyright 2020 by the American Psychological Association.

Credits for sample student paper templates

These papers came from real students who gave their permission to have them edited and posted by APA.

U.S. flag

An official website of the United States government

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

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

  • Publications
  • Account settings

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

  • Advanced Search
  • Journal List
  • v.25(3); 2014 Oct

Logo of ejifcc

How to Write a Scientific Paper: Practical Guidelines

Edgard delvin.

1 Centre de recherche, CHU Sainte-Justine

2 Département de Biochimie, Université de Montréal, Montréal, Canada

Tahir S. Pillay

3 Department of Chemical Pathology, Faculty of Health Sciences, University of Pretoria

4 Division of Chemical Pathology, University of Cape Town

5 National Health Laboratory Service, CTshwane Academic Division, Pretoria, South Africa

Anthony Newman

6 Life Sciences Department, Elsevier, Amsterdam, The Netherlands

Precise, accurate and clear writing is essential for communicating in health sciences, as publication is an important component in the university criteria for academic promotion and in obtaining funding to support research. In spite of this, the development of writing skills is a subject infrequently included in the curricula of faculties of medicine and allied health sciences. Therefore clinical investigators require tools to fill this gap. The present paper presents a brief historical background to medical publication and practical guidelines for writing scientific papers for acceptance in good journals.

INTRODUCTION

A scientific paper is the formal lasting record of a research process. It is meant to document research protocols, methods, results and conclusions derived from an initial working hypothesis. The first medical accounts date back to antiquity. Imhotep, Pharaoh of the 3 rd Dynasty, could be considered the founder of ancient Egyptian medicine as he has been credited with being the original author of what is now known as the Edwin Smith Papyrus ( Figure 1 ). The Papyrus, by giving some details on cures and anatomical observations, sets the basis of the examination, diagnosis, treatment, and prognosis of numerous diseases. Closer to the Common Era, in 460 BCE, Hippocrates wrote 70 books on medicine. In 1020, the Golden age of the Muslim Culture, Ibn Sina, known as Avicenna ( Figure 2a ), recorded the Canon of medicine that was to become the most used medical text in Europe and Middle East for almost half a millennium. This was followed in the beginning of the 12 th Century bytheextensivetreatiseofMaimonides( Figure 2b ) (Moses ben Maimon) on Greek and Middle Eastern medicine. Of interest, by the end of the 11 th Century Trotula di Ruggiero, a woman physician, wrote several influential books on women’s ailment. A number of other hallmark treatises also became more accessible, thanks to the introduction of the printing press that allowed standardization of the texts. One example is the De Humani Corporis Fabrica by Vesalius which contains hundreds of illustrations of human dissection. Thomas A Lang provides an excellent concise history of scientific publications [ 1 ]. These were the days when writing and publishing scientific or philosophical works were the privilege of the few and hence there was no or little competition and no recorded peer reviewing system. Times have however changed, and contemporary scientists have to compose with an increasingly harsh competition in attracting editors and publishers attention. As an example, the number of reports and reviews on obesity and diabetes has increased from 400 to close to 4000/year and 50 to 600/year respectively over a period of 20 years ( Figure 3 ). The present article, essentially based on TA Lang’s guide for writing a scientific paper [ 1 ], will summarize the steps involved in the process of writing a scientific report and in increasing the likelihood of its acceptance.

This manuscript, written in 1600 BCE, is regarded as a copy of several earlier works ( 3000 BCE). It is part of a textbook on surgery the examination, diagnosis, treatment, and prognosis of numerous ailments. BCE: Before the Common Era.

The Edwin Smith Papyrus (≈3000 BCE)

Figure 2a Avicenna 973-1037 C.E.Figure 2b Maimonides, 1135-1204 C.E.

Avicenna and Maimonides

Orange columns: original research papers; Green columns: reviews

Annual publication load in the field of obesity and diabetes over 20 years.

Reasons for publishing are varied. One may write to achieve a post-graduate degree, to obtain funding for pursuing research or for academic promotion. While all 3 reasons are perfectly legitimate, one must ask whether they are sufficient to be considered by editors, publishers and reviewers. Why then should the scientist write? The main reason is to provide to the scientific community data based on hypotheses that are innovative and thus to advance the understanding in a specific domain. One word of caution however, is that if a set of experiments has not been done or reported, it does not mean that it should be. It may simply reflect a lack of interest in it.

DECIDING ON PUBLISHING AND TARGETING THE JOURNAL

In order to assist with the decision process, pres-ent your work orally first to colleagues in your field who may be more experienced in publishing. This step will help you in gauging whether your work is publishable and in shaping the paper.

Targeting the journal, in which you want to present your data, is also a critical step and should be done before starting to write. One hint is to look for journals that have published similar work to yours, and that aims readers most likely to be interested in your research. This will allow your article to be well read and cited. These journals are also those that you are most likely to read on a regular basis and to cite abundantly. The next step is to decide whether you submit your manuscript to a top-ranking impact factor journal or to a journal of lower prestige. Although it is tempting to test the waters, or to obtain reviewers comments, be realistic about the contribution your work provides and submit to a journal with an appropriate rank.

Do not forget that each rejection delays publication and that the basin of reviewers within your specialty is shallow. Thus repeated submission to different journals could likely result in having your work submitted for review to the same re-viewer.

DECIDING ON THE TYPE OF MANUSCRIPT

There are several types of scientific reports: observational, experimental, methodological, theoretical and review. Observational studies include 1) single-case report, 2) collective case reports on a series of patients having for example common signs and symptoms or being followed-up with similar protocols, 3) cross-sectional, 4) cohort studies, and 5) case-control studies. The latter 3 could be perceived as epidemiological studies as they may help establishing the prevalence of a condition, and identify a defined population with and without a particular condition (disease, injury, surgical complication). Experimental reports deal with research that tests a research hypothesis through an established protocol, and, in the case of health sciences, formulate plausible explanations for changes in biological systems. Methodological reports address for example advances in analytical technology, statistical methods and diagnostic approach. Theoretical reports suggest new working hypotheses and principles that have to be supported or disproved through experimental protocols. The review category can be sub-classified as narrative, systematic and meta-analytic. Narrative reviews are often broad overviews that could be biased as they are based on the personal experience of an expert relying on articles of his or her own choice. Systematic reviews and meta-analyses are based on reproducible procedures and on high quality data. Researchers systematically identify and analyze all data collected in articles that test the same working hypothesis, avoiding selection bias, and report the data in a systematic fashion. They are particularly helpful in asking important questions in the field of healthcare and are often the initial step for innovative research. Rules or guidelines in writing such report must be followed if a quality systematic review is to be published.

For clinical research trials and systematic reviews or meta-analyses, use the Consort Statement (Consolidated Standards Of Reporting Trials) and the PRISMA Statement (Preferred Reporting Items for Systematic reviews and Meta-Analyses) respectively [ 2 , 3 ]. This assures the editors and the reviewers that essential elements of the trials and of the reviews were tackled. It also speeds the peer review process. There are several other Statements that apply to epidemiological studies [ 4 ], non-randomized clinical trials [ 5 ], diagnostic test development ( 6 ) and genetic association studies ( 7 ). The Consortium of Laboratory Medicine Journal Editors has also published guidelines for reporting industry-sponsored laboratory research ( 8 ).

INITIAL STEPS IN THE PROCESS OF WRITING A SCIENTIFIC DOCUMENT

Literature review is the initial and essential step before starting your study and writing the scientific report based on it. In this process use multiple databases, multiple keyword combinations. It will allow you to track the latest development in your field and thus avoid you to find out that someone else has performed the study before you, and hence decrease the originality of your study. Do not forget that high-ranking research journals publish results of enough importance and interest to merit their publication.

Determining the authorship and the order of authorship, an ethical issue, is the second essential step, and is unfortunately often neglected. This step may avoid later conflicts as, despite existing guidelines, it remains a sensitive issue owing to personal biases and the internal politics of institutions. The International Committee of Medical Editors has adopted the following guidelines for the biomedical sciences ( 9 ).

“Authorship credit should be based only on: 1) Substantial contributions to the conception and design, or acquisition of data, or analysis and interpretation of data; 2) Drafting the article or revising it critically for important intellectual content; and 3) Final approval of the version to be published. Conditions 1, 2 and 3 must be all met. Acquisition of funding, the collections of data, or general supervision of the research group, by themselves, do not justify authorship.” ( 9 , 10 )

The order of authorship should reflect the individual contribution to the research and to the publication, from most to least ( 11 ). The first author usually carries out the lead for the project reported. However the last author is often mistakenly perceived as the senior author. This is perpetuated from the European tradition and is discouraged. As there are divergent conventions among journals, the order of authorship order may or may not reflect the individual contributions; with the exception that the first author should be the one most responsible for the work.

WRITING EFFECTIVELY

Effective writing requires that the text helps the readers 1) understand the content and the context, 2) remember what the salient points are, 3) find the information rapidly and, 4) use or apply the information given. These cardinal qualities should be adorned with the precise usage of the language, clarity of the text, inclu-siveness of the information, and conciseness. Effective writing also means that you have to focus on the potential readers’ needs. Readers in science are informed individuals who are not passive, and who will formulate their own opinion of your writing whether or not the meaning is clear. Therefore you need to know who your audience is. The following 4 questions should help you writing a reader-based text, meaning written to meet the information needs of readers [ 12 ].

What do you assume your readers already know? In other words, which terms and concepts can you use without explanation, and which do you have to define?

What do they want to know? Readers in science will read only if they think they will learn something of value.

What do they need to know? Your text must contain all the information necessary for the reader to understand it, even if you think this information id obvious to them.

What do they think they know that is not so? Correcting misconceptions can be an important function of communication, and persuading readers to change their minds can be a challenging task.

WRITING THE SCIENTIFIC PAPER

Babbs and Tacker ’ s advice to write as much of the paper before performing the research project or experimental protocol may, at first sight, seem unexpected and counterintuitive [ 13 ], but in fact it is exactly what is being done when writing a research grant application. It will allow you to define the authorship alluded to before. The following section will briefly review the structure of the different sections of a manuscript and describe their purpose.

Reading the instructions to authors of the Journal you have decided to submit your manuscript is the first important step. They provide you with the specific requirements such as the way of listing the authors, type of abstract, word, figure or table limits and citation style. The Mulford Library of University of Toledo website contains instructions to authors for over 3000 journals ( http://mulford.meduoiho.edu/instr/ ).

The general organization of an article follows the IMRAD format (Introduction, Methods, Results, and Discussion). These may however vary. For instance, in clinical research or epidemiology studies, the methods section will include details on the subjects included, and there will be a statement of the limitation of the study. Although conclusions may not always be part of the structure, we believe that it should, even in methodological reports.

The tile page provides essential information so that the editor, reviewers, and readers will identify the manuscript and the authors at a glance as well as enabling them to classify the field to which the article pertains.

The title page must contain the following:

  • The tile of the article – it is an important part of the manuscript as it is the most often read and will induce the interested readers to pursue further. Therefore the title should be precise, accurate, specific and truthful;
  • Each author’s given name (it may be the full name or initials) and family name;
  • Each author’s affiliation;
  • Some journals ask for highest academic degree;
  • A running title that is usually limited to a number of characters. It must relate to the full title;
  • Key words that will serve for indexing;
  • For clinical studies, the trial’s registration number;
  • The name of the corresponding author with full contact information.

The abstract is also an important section of your manuscript. Importantly, the abstract is the part of the article that your peers will see when consulting publication databases such as PubMed. It is the advertisement to your work and will strongly influence the editor deciding whether it will be submitted to reviewers or not. It will also help the readers decide to read the full article. Hence it has to be comprehensible on its own. Writing an abstract is challenging. You have to carefully select the content and, while being concise, assure to deliver the essence of your manuscript.

Without going into details, there are 3 types of abstracts: descriptive, informative and structured. The descriptive abstract is particularly used for theoretical, methodological or review articles. It usually consists of a single paragraph of 150 words or less. The informative abstract, the most common one, contains specific information given in the article and, are organized with an introduction (background, objectives), methods, results and discussion with or without conclusion. They usually are 150 to 250 words in length. The structured abstract is in essence an informative abstract with sections labeled with headings. They may also be longer and are limited to 250 to 300 words. Recent technology also allows for graphical or even video abstracts. The latter are interesting in the context of cell biology as they enable the investigator to illustrate ex vivo experiment results (phagocytosis process for example).

Qualities of abstracts:

  • Understood without reading the full paper. Shoul dcontain no abbreviations.lf abbreviations are used, they must be defined. This however removes space for more important information;
  • Contains information consistent with the full report. Conclusions in the abstract must match those given in the full report;
  • Is attractive and contains information needed to decide whether to read the full report.

Introduction

The introduction has 3 main goals: to establish the need and importance of your research, to indicate how you have filled the knowledge gap in your field and to give your readers a hint of what they will learn when reading your paper. To fulfil these goals, a four-part introduction consisting of a background statement, a problem statement, an activity statement and a forecasting statement, is best suited. Poorly defined background information and problem setting are the 2 most common weaknesses encountered in introductions. They stem from the false perception that peer readers know what the issue is and why the study to solve it is necessary. Although not a strict rule, the introduction in clinical science journals should target only references needed to establish the rationale for the study and the research protocol. This differ from more basic science or cell biology journals, for which a longer and elaborate introduction may be justified because the research at hand consists of several approaches each requiring background and justification.

The 4-part introduction consists of:

  • A background statement that provides the context and the approach of the research;
  • A problem statement that describes the nature, scope and importance of the problem or the knowledge gap;
  • An activity statement, that details the research question, sets the hypothesis and actions undertaken for the investigation;
  • A forecasting statement telling the readers whattheywillfìndwhen readingyourarticle [ 14 ].

Methods section

This section may be named “Materials and Methods”, “Experimental section” or “Patients and Methods” depending upon the type of journal. Its purpose to allow your readers to provide enough information on the methods used for your research and to judge on their adequacy. Although clinical and “basic” research protocols differ, the principles involved in describing the methods share similar features. Hence, the breadth of what is being studied and how the study can be performed is common to both. What differ are the specific settings. For example, when a study is conducted on humans, you must provide, up front, assurance that it has received the approval of you Institution Ethics Review Board (IRB) and that participants have provided full and informed consent. Similarly when the study involves animals, you must affirm that you have the agreement from your Institutional Animal Care and Use Committee (IACUC). These are too often forgotten, and Journals (most of them) abiding to the rules of the Committee on Publication Ethics (COPE) and World Association of Medical Editors (WAME) will require such statement. Although journals publishing research reports in more fundamental science may not require such assurance, they do however also follow to strict ethics rules related to scientific misconduct or fraud such as data fabrication, data falsification. For clinical research papers, you have to provide information on how the participants were selected, identify the possible sources of bias and confounding factors and how they were diminished.

In terms of the measurements, you have to clearly identify the materials used as well as the suppliers with their location. You should also be unambiguous when describing the analytical method. If the method has already been published, give a brief account and refer to the original publication (not a review in which the method is mentioned without a description). If you have modified it, you have to provide a detailed account of the modifications and you have to validate its accuracy, precision and repeatability. Mention the units in which results are reported and, if necessary, include the conversion factors [mass units versus “système international” (S.I.)]. In clinical research, surrogate end-points are often used as biomarkers. Under those circumstances, you must show their validity or refer to a study that has already shown that are valid.

In cases of clinical trials, the Methods section should include the study design, the patient selection mode, interventions, type of outcomes.

Statistics are important in assuring the quality of the research project. Hence, you should consult a biostatistician at the time of devising the research protocol and not after having performed the experiments or the clinical trial.

The components of the section on statistics should include:

  • The way the data will be reported (mean, median, centiles for continuous data);
  • Details on participant assignments to the different groups (random allocation, consecutive entry);
  • Statistical comparison tools (parametric or non parametric statistics, paired or unpaired t-tests for normally distributed data and so on);
  • The statistical power calculation when determining the sample size to obtain valid and significant comparisons together with the a level;
  • The statistical software package used in the analysis.

Results section

The main purpose of the results section is to report the data that were collected and their relationship. It should also provide information on the modifications that have taken place because of unforeseen events leading to a modification of the initial protocol (loss of participants, reagent substitution, loss of data).

  • Report results as tables and figures whenever possible, avoid duplication in the text. The text should summarize the findings;
  • Report the data with the appropriate descriptive statistics;
  • Report any unanticipated events that could affect the results;
  • Report a complete account of observations and explanations for missing data (patient lost).

The discussion should set your research in context, reinforce its importance and show how your results have contributed to the further understanding of the problem posed. This should appear in the concluding remarks. The following organization could be helpful.

  • Briefly summarize the main results of your study in one or two paragraphs, and how they support your working hypothesis;
  • Provide an interpretation of your results and show how they logically fit in an overall scheme (biological or clinical);
  • Describe how your results compare with those of other investigators, explain the differences observed;
  • Discuss how your results may lead to a new hypothesis and further experimentation, or how they could enhance the diagnostic procedures.
  • Provide the limitations of your study and steps taken to reduce them. This could be placed in the concluding remarks.

Acknowledgements

The acknowledgements are important as they identify and thank the contributors to the study, who do not meet the criteria as co-authors. They also include the recognition of the granting agency. In this case the grant award number and source is usually included.

Declaration of competing interests

Competing interests arise when the author has more than one role that may lead to a situation where there is a conflict of interest. This is observed when the investigator has a simultaneous industrial consulting and academic position. In that case the results may not be agreeable to the industrial sponsor, who may impose a veto on publication or strongly suggest modifications to the conclusions. The investigator must clear this issue before starting the contracted research. In addition, the investigator may own shares or stock in the company whose product forms the basis of the study. Such conflicts of interest must be declared so that they are apparent to the readers.

Acknowledgments

The authors thank Thomas A Lang, for his advice in the preparation of this manuscript.

Deakin University Australia

  • Deakin Home
  • Student news
  • StudentConnect
  • Help and contact us

Writing a scientific report

Before you begin writing a scientific paper, it is useful to know how to read one. Being familiar with the language and structure will not only improve your reading skills but it will help you to construct your own reports more effectively.

Reading a scientific paper

I think the important thing in reading a scientific paper is understanding the structure. Once you understand the structure, it’s easy to select information because you know where to look. - Dennis Farrugia Language and Learning Adviser.

Choose credible and reliable sources

Before your first reading of a scientific paper, check that the information you are reading is  credible, reliable and accurate .

Check with your tutor if you need further assistance with choosing appropriate sources. Also, explore the Deakin Library's  library guides in your discipline.

Reading and taking notes

Identifying the different parts of a scientific text can help you read more effectively and find relevant information for your assignments. In this video, a language and learning adviser discusses the importance of understanding structure when reading a scientific paper.

Pay attention to the different sections commonly used in a journal article to better understand the research being presented. You can then use your knowledge of these sections to scan papers and to extract key information for your own notes. The different sections include:

  • Abstract - overview of paper.
  • Introduction - background, context (resolving a problem or gap in the research), the aims of the research.
  • Methodology – how the research was conducted, how the data was collected and analysed, sample sizes.
  • Results  or  Findings - results only (no discussion).
  • Discussion - discussion of results and how it relates to other research and the implications of this.
  • Conclusion - explores implications for further research.

After you have gained some understanding of the research, start thinking critically about the information by:

  • comparing this research to other studies on the same topic
  • judging whether there was enough data to support the claims being made
  • suggesting what the limitations of the data might mean for the results.

Reading and analysing scientific texts (PDF, 70.0KB)

Guide to writing lab and field reports (PDF, 91.0KB)

Writing strategies

Scientific writing is very precise, so it’s important to make sure you’re as concise and clear as possible. Being clear with your purpose helps you stay focused on what you’re writing about. - Dennis Farrugia Language and Learning Adviser

In this video, a language and learning adviser provides some useful language tips for writing a scientific paper. In summary, these tips are:

  • Be clear about the purpose of the paper.
  • Use precise language.
  • Be aware of your use of verb tense (past tense is often used, as you are reporting on past events in the lab/field).
  • Use  cautious/tentative and objective language.
  • Order your ideas logically, using the appropriate structure.

Defining terms

An important feature of science writing is knowing how to define terms. As you read, take note of how scientific terms are explained and categorised. Once you have identified some of these models of writing you can then apply them in your own work.

Type of definitionExample
Initial definition
Defining sub-classes
Commenting on a system of classification



Referring to other authors' definitions
Defining specific terms that are used in your own report

For more information on using tentative language, classifying, listing and reporting results, visit the  Manchester academic phrasebank .

Download the  guide to writing lab and field reports (PDF) for further examples of the characteristics of scientific writing.

When you are asked to write a report on investigations you carry out in labs or when you go on fieldwork, it is important to recognise that these reports are structured differently from other types of research reports and essays.

Lab or fieldwork reports are based on detailed observations of the aims, methods and procedures of your experiments or fieldwork investigations, so it is important to keep very precise and detailed notes when you are out in the field or working in the lab.

Download the  Guide to writing lab and field reports (PDF, 91.0KB) on this page for an overview of the structure of reports, as well as some language tips for each section of the report.

Note: Always follow the assessment instructions provided in your unit. This guide provides general advice only.

You might also like:

  • Evaluating sources
  • Academic style
  • Critical thinking, reading and note taking

Book an appointment with a language and learning adviser; or First Nations students can book with the First Nations academic support links

  • Book an appointment
  • First Nations Academic Support bookings
  • Email First Nations Academic Support

We use cookies to improve your experience. You consent to the use of our cookies if you proceed. Visit our Privacy policy for more information.

Writing an Introduction for a Scientific Paper

Dr. michelle harris, dr. janet batzli, biocore.

This section provides guidelines on how to construct a solid introduction to a scientific paper including background information, study question , biological rationale, hypothesis , and general approach . If the Introduction is done well, there should be no question in the reader’s mind why and on what basis you have posed a specific hypothesis.

Broad Question : based on an initial observation (e.g., “I see a lot of guppies close to the shore. Do guppies like living in shallow water?”). This observation of the natural world may inspire you to investigate background literature or your observation could be based on previous research by others or your own pilot study. Broad questions are not always included in your written text, but are essential for establishing the direction of your research.

Background Information : key issues, concepts, terminology, and definitions needed to understand the biological rationale for the experiment. It often includes a summary of findings from previous, relevant studies. Remember to cite references, be concise, and only include relevant information given your audience and your experimental design. Concisely summarized background information leads to the identification of specific scientific knowledge gaps that still exist. (e.g., “No studies to date have examined whether guppies do indeed spend more time in shallow water.”)

Testable Question : these questions are much more focused than the initial broad question, are specific to the knowledge gap identified, and can be addressed with data. (e.g., “Do guppies spend different amounts of time in water <1 meter deep as compared to their time in water that is >1 meter deep?”)

Biological Rationale : describes the purpose of your experiment distilling what is known and what is not known that defines the knowledge gap that you are addressing. The “BR” provides the logic for your hypothesis and experimental approach, describing the biological mechanism and assumptions that explain why your hypothesis should be true.

The biological rationale is based on your interpretation of the scientific literature, your personal observations, and the underlying assumptions you are making about how you think the system works. If you have written your biological rationale, your reader should see your hypothesis in your introduction section and say to themselves, “Of course, this hypothesis seems very logical based on the rationale presented.”

  • A thorough rationale defines your assumptions about the system that have not been revealed in scientific literature or from previous systematic observation. These assumptions drive the direction of your specific hypothesis or general predictions.
  • Defining the rationale is probably the most critical task for a writer, as it tells your reader why your research is biologically meaningful. It may help to think about the rationale as an answer to the questions— how is this investigation related to what we know, what assumptions am I making about what we don’t yet know, AND how will this experiment add to our knowledge? *There may or may not be broader implications for your study; be careful not to overstate these (see note on social justifications below).
  • Expect to spend time and mental effort on this. You may have to do considerable digging into the scientific literature to define how your experiment fits into what is already known and why it is relevant to pursue.
  • Be open to the possibility that as you work with and think about your data, you may develop a deeper, more accurate understanding of the experimental system. You may find the original rationale needs to be revised to reflect your new, more sophisticated understanding.
  • As you progress through Biocore and upper level biology courses, your rationale should become more focused and matched with the level of study e ., cellular, biochemical, or physiological mechanisms that underlie the rationale. Achieving this type of understanding takes effort, but it will lead to better communication of your science.

***Special note on avoiding social justifications: You should not overemphasize the relevance of your experiment and the possible connections to large-scale processes. Be realistic and logical —do not overgeneralize or state grand implications that are not sensible given the structure of your experimental system. Not all science is easily applied to improving the human condition. Performing an investigation just for the sake of adding to our scientific knowledge (“pure or basic science”) is just as important as applied science. In fact, basic science often provides the foundation for applied studies.

Hypothesis / Predictions : specific prediction(s) that you will test during your experiment. For manipulative experiments, the hypothesis should include the independent variable (what you manipulate), the dependent variable(s) (what you measure), the organism or system , the direction of your results, and comparison to be made.

We hypothesized that reared in warm water will have a greater sexual mating response.

(The dependent variable “sexual response” has not been defined enough to be able to make this hypothesis testable or falsifiable. In addition, no comparison has been specified— greater sexual mating response as compared to what?)

We hypothesized that ) reared in warm water temperatures ranging from 25-28 °C ( ) would produce greater ( ) numbers of male offspring and females carrying haploid egg sacs ( ) than reared in cooler water temperatures of 18-22°C.

If you are doing a systematic observation , your hypothesis presents a variable or set of variables that you predict are important for helping you characterize the system as a whole, or predict differences between components/areas of the system that help you explain how the system functions or changes over time.

We hypothesize that the frequency and extent of algal blooms in Lake Mendota over the last 10 years causes fish kills and imposes a human health risk.

(The variables “frequency and extent of algal blooms,” “fish kills” and “human health risk” have not been defined enough to be able to make this hypothesis testable or falsifiable. How do you measure algal blooms? Although implied, hypothesis should express predicted direction of expected results [ , higher frequency associated with greater kills]. Note that cause and effect cannot be implied without a controlled, manipulative experiment.)

We hypothesize that increasing ( ) cell densities of algae ( ) in Lake Mendota over the last 10 years is correlated with 1. increased numbers of dead fish ( ) washed up on Madison beaches and 2. increased numbers of reported hospital/clinical visits ( .) following full-body exposure to lake water.

Experimental Approach : Briefly gives the reader a general sense of the experiment, the type of data it will yield, and the kind of conclusions you expect to obtain from the data. Do not confuse the experimental approach with the experimental protocol . The experimental protocol consists of the detailed step-by-step procedures and techniques used during the experiment that are to be reported in the Methods and Materials section.

Some Final Tips on Writing an Introduction

  • As you progress through the Biocore sequence, for instance, from organismal level of Biocore 301/302 to the cellular level in Biocore 303/304, we expect the contents of your “Introduction” paragraphs to reflect the level of your coursework and previous writing experience. For example, in Biocore 304 (Cell Biology Lab) biological rationale should draw upon assumptions we are making about cellular and biochemical processes.
  • Be Concise yet Specific: Remember to be concise and only include relevant information given your audience and your experimental design. As you write, keep asking, “Is this necessary information or is this irrelevant detail?” For example, if you are writing a paper claiming that a certain compound is a competitive inhibitor to the enzyme alkaline phosphatase and acts by binding to the active site, you need to explain (briefly) Michaelis-Menton kinetics and the meaning and significance of Km and Vmax. This explanation is not necessary if you are reporting the dependence of enzyme activity on pH because you do not need to measure Km and Vmax to get an estimate of enzyme activity.
  • Another example: if you are writing a paper reporting an increase in Daphnia magna heart rate upon exposure to caffeine you need not describe the reproductive cycle of magna unless it is germane to your results and discussion. Be specific and concrete, especially when making introductory or summary statements.

Where Do You Discuss Pilot Studies? Many times it is important to do pilot studies to help you get familiar with your experimental system or to improve your experimental design. If your pilot study influences your biological rationale or hypothesis, you need to describe it in your Introduction. If your pilot study simply informs the logistics or techniques, but does not influence your rationale, then the description of your pilot study belongs in the Materials and Methods section.  

from an Intro Ecology Lab:

         Researchers studying global warming predict an increase in average global temperature of 1.3°C in the next 10 years (Seetwo 2003). are small zooplankton that live in freshwater inland lakes. They are filter-feeding crustaceans with a transparent exoskeleton that allows easy observation of heart rate and digestive function. Thomas et al (2001) found that heart rate increases significantly in higher water temperatures are also thought to switch their mode of reproduction from asexual to sexual in response to extreme temperatures. Gender is not mediated by genetics, but by the environment. Therefore, reproduction may be sensitive to increased temperatures resulting from global warming (maybe a question?) and may serve as a good environmental indicator for global climate change.

         In this experiment we hypothesized that reared in warm water will switch from an asexual to a sexual mode of reproduction. In order to prove this hypothesis correct we observed grown in warm and cold water and counted the number of males observed after 10 days.

Comments:

Background information

·       Good to recognize as a model organism from which some general conclusions can be made about the quality of the environment; however no attempt is made to connect increased lake temperatures and gender. Link early on to increase focus.

·       Connection to global warming is too far-reaching. First sentence gives impression that Global Warming is topic for this paper. Changes associated with global warming are not well known and therefore little can be concluded about use of as indicator species.

·       Information about heart rate is unnecessary because heart rate in not being tested in this experiment.

Rationale

·       Rationale is missing; how is this study related to what we know about D. magna survivorship and reproduction as related to water temperature, and how will this experiment contribute to our knowledge of the system?

·       Think about the ecosystem in which this organism lives and the context. Under what conditions would D. magna be in a body of water with elevated temperatures?

Hypothesis

·       Not falsifiable; variables need to be better defined (state temperatures or range tested rather than “warm” or “cold”) and predict direction and magnitude of change in number of males after 10 days.

·       It is unclear what comparison will be made or what the control is

·       What dependent variable will be measured to determine “switch” in mode of reproduction (what criteria are definitive for switch?)

Approach

·       Hypotheses cannot be “proven” correct. They are either supported or rejected.

Introduction

         are small zooplankton found in freshwater inland lakes and are thought to switch their mode of reproduction from asexual to sexual in response to extreme temperatures (Mitchell 1999). Lakes containing have an average summer surface temperature of 20°C (Harper 1995) but may increase by more than 15% when expose to warm water effluent from power plants, paper mills, and chemical industry (Baker et al. 2000). Could an increase in lake temperature caused by industrial thermal pollution affect the survivorship and reproduction of ?

         The sex of is mediated by the environment rather than genetics. Under optimal environmental conditions, populations consist of asexually reproducing females. When the environment shifts may be queued to reproduce sexually resulting in the production of male offspring and females carrying haploid eggs in sacs called ephippia (Mitchell 1999).

         The purpose of this laboratory study is to examine the effects of increased water temperature on survivorship and reproduction. This study will help us characterize the magnitude of environmental change required to induce the onset of the sexual life cycle in . Because are known to be a sensitive environmental indicator species (Baker et al. 2000) and share similar structural and physiological features with many aquatic species, they serve as a good model for examining the effects of increasing water temperature on reproduction in a variety of aquatic invertebrates.

         We hypothesized that populations reared in water temperatures ranging from 24-26 °C would have lower survivorship, higher male/female ratio among the offspring, and more female offspring carrying ephippia as compared with grown in water temperatures of 20-22°C. To test this hypothesis we reared populations in tanks containing water at either 24 +/- 2°C or 20 +/- 2°C. Over 10 days, we monitored survivorship, determined the sex of the offspring, and counted the number of female offspring containing ephippia.

Comments:

Background information

·       Opening paragraph provides good focus immediately. The study organism, gender switching response, and temperature influence are mentioned in the first sentence. Although it does a good job documenting average lake water temperature and changes due to industrial run-off, it fails to make an argument that the 15% increase in lake temperature could be considered “extreme” temperature change.

·       The study question is nicely embedded within relevant, well-cited background information. Alternatively, it could be stated as the first sentence in the introduction, or after all background information has been discussed before the hypothesis.

Rationale

·       Good. Well-defined purpose for study; to examine the degree of environmental change necessary to induce the Daphnia sexual life
cycle.

How will introductions be evaluated? The following is part of the rubric we will be using to evaluate your papers.

 

0 = inadequate

(C, D or F)

1 = adequate

(BC)

2 = good

(B)

3 = very good

(AB)

4 = excellent

(A)

Introduction

BIG PICTURE: Did the Intro convey why experiment was performed and what it was designed to test?

 

Introduction provides little to no relevant information. (This often results in a hypothesis that “comes out of nowhere.”)

Many key components are very weak or missing; those stated are unclear and/or are not stated concisely. Weak/missing components make it difficult to follow the rest of the paper.

e.g., background information is not focused on a specific question and minimal biological rationale is presented such that hypothesis isn’t entirely logical

 

Covers most key components but could be done much more logically, clearly, and/or concisely.

e.g., biological rationale not fully developed but still supports hypothesis. Remaining components are done reasonably well, though there is still room for improvement.

Concisely & clearly covers all but one key component (w/ exception of rationale; see left) clearly covers all key components but could be a little more concise and/or clear.

e.g., has done a reasonably nice job with the Intro but fails to state the approach OR has done a nice job with Intro but has also included some irrelevant background information

 

Clearly, concisely, & logically presents all key components: relevant & correctly cited background information, question, biological rationale, hypothesis, approach.

This document originally came from the Journal of Mammalogy courtesy of Dr. Ronald Barry, a former editor of the journal.

Banner

  • Scientific Lab Reports
  • Understanding the Assignment
  • Need a Topic?
  • Evaluating Sources
  • Brainstorming Strategies
  • Drafting Strategies
  • Thesis Formulation
  • Introductions
  • Conclusions
  • Show Don't Tell
  • Expand Your Draft
  • Flow & Lexical Coherence
  • Revision Checklist
  • Introduction to Style and Grammar
  • Apostrophes
  • Article Usage for ESL Learners
  • Capitalization
  • Clarity: Get Rid of Nominalizations
  • Cohesion: Does my Paragraph Flow?
  • Commas and Colons
  • Conciseness
  • Confusing Words
  • Parallel Structure
  • Passive Voice
  • Quotation Marks
  • Run-on Sentences
  • Subject-Verb Agreement
  • Writing Mechanics
  • Other Styles
  • Art History This link opens in a new window
  • Programming Lab Reports

Writing a Lab Report

Link to other resources.

  • Screenwriting
  • Publication Opportunities
  • Meet with a Tutor This link opens in a new window

Academic Resource Center Hours :

Monday-Thursday: 8:00 AM-6:00 PM

Friday: 8:00 AM-4:00 PM

Phone Number : 310-338-2847

Email :  [email protected]

www.lmu.edu/arc

We are located in Daum Hall on the 2nd floor!

Writing a scientific lab report is significantly different from writing for other classes like philosophy, English, and history. The most prominent form of writing in biology, chemistry, and environmental science is the lab report, which is a formally written description of results and discoveries found in an experiment. College lab reports should emulate and follow the same formats as reports found in scholarly journals, such as Nature , Cell , and The American Journal of Biochemistry .

Report Format

Title: The title says what you did. It should be brief (aim for ten words or less) and describe the main point of the experiment or investigation.

  • Example:  Caffeine Increases Amylase Activity in the Mealworm ( Tenebrio molitar).
  • If you can, begin your title using a keyword rather than an article like “The” or “A.”

Abstract: An abstract is a very concise summary of the purpose of the report, data presented, and major conclusions in about 100 - 200 words.  Abstracts are also commonly required for conference/presentation submissions because they summarize all of the essential materials necessary to understand the purpose of the experiment. They should consist of a background sentence , an introduction sentence , your hypothesis/purpose of the experiment, and a sentence about the results and what this means.

Introduction: The introduction of a lab report defines the subject of the report, provides background information and relevant studies, and outlines scientific purpose(s) and/or objective(s).

  • The introduction is a place to provide the reader with necessary research on the topic and properly cite sources used.
  • Summarizes the current literature on the topic including primary and secondary sources.
  • Introduces the paper’s aims and scope.
  • States the purpose of the experiment and the hypothesis.

Materials and Methods: The materials and methods section is a vital component of any formal lab report. This section of the report gives a detailed account of the procedure that was followed in completing the experiment as well as all important materials used. (This includes bacterial strains and species names in tests using living subjects.)

  • Discusses the procedure of the experiment in as much detail as possible.
  • Provides information about participants, apparatus, tools, substances, location of experiment, etc.
  • For field studies, be sure to clearly explain where and when the work was done.
  • It must be written so that anyone can use the methods section as instructions for exact replications.
  • Don’t hesitate to use subheadings to organize these categories.
  • Practice proper scientific writing forms. Be sure to use the proper abbreviations for units. Example: The 50mL sample was placed in a 5ºC room for 48hrs.

Results: The results section focuses on the findings, or data, in the experiment, as well as any statistical tests used to determine their significance.

  • Concentrate on general trends and differences and not on trivial details.
  • Summarize the data from the experiments without discussing their implications (This is where all the statistical analyses goes.)
  • Organize data into tables, figures, graphs, photographs, etc.  Data in a table should not be duplicated in a graph or figure. Be sure to refer to tables and graphs in the written portion, for example, “Figure 1 shows that the activity....”
  • Number and title all figures and tables separately, for example, Figure 1 and Table 1 and include a legend explaining symbols and abbreviations. Figures and graphs are labeled below the image while tables are labeled above.

  Discussion: The discussion section interprets the results, tying them back to background information and experiments performed by others in the past.This is also the area where further research opportunities shold be explored.

  • Interpret the data; do not restate the results.
  • Observations should also be noted in this section, especially anything unusual which may affect your results.

For example, if your bacteria was incubated at the wrong temperature or a piece of equipment failed mid-experiment, these should be noted in the results section.

  • Relate results to existing theories and knowledge.This can tie back to your introduction section because of the background you provided.
  • Explain the logic that allows you to accept or reject your original hypotheses.
  • Include suggestions for improving your techniques or design, or clarify areas of doubt for further research.

Acknowledgements and References: A references list should be compiled at the end of the report citing any works that were used to support the paper. Additionally, an acknowledgements section should be included to acknowledge research advisors/ partners, any group or person providing funding for the research and anyone outside the authors who contributed to the paper or research.

General Tips

  • In scientific papers, passive voice is perfectly acceptable. On the other hand, using “I” or “we” is not.

          Incorrect: We found that caffeine increased amylase levels in Tenebrio molitar.  Correct: It was discovered that caffeine increased amylase levels in Tenebrio molitar.   

  • It is expected that you use as much formal (bland) language and scientific terminology as you can. There should be no emphasis placed on “expressing yourself” or “keeping it interesting”; a lab report is not a narrative.
  • In a lab report, it is important to get to the point. Be descriptive enough that your audience can understand the experiment, but strive to be concise.
  • << Previous: Programming Lab Reports
  • Next: Screenwriting >>
  • Last Updated: Aug 27, 2024 9:44 AM
  • URL: https://libguides.lmu.edu/writing

Science Essay Examples

Caleb S.

Best Science Essay Examples to Learn From

Published on: May 3, 2023

Last updated on: Jan 31, 2024

Science Essay Examples

Share this article

Are you struggling to write a science essay that stands out? 

Are you tired of feeling overwhelmed by scientific jargon and complicated concepts? 

You're not alone. 

Science essays can be a challenge for even the most dedicated students. It's no wonder that so many students struggle to produce top-notch papers.

But fear not! 

In this blog post, we'll provide you with some science essay examples and tips. We will help you write a top-notch paper that impresses your professor and earns you a high grade. 

So buckle up and get ready to tackle science essays like a pro!

On This Page On This Page -->

Science Essay Examples for Students

Writing a science essay can be a daunting task for students. However, with the right guidance and examples, it can also be a rewarding and enlightening experience.

Here, we'll provide you with examples so you can elevate your own writing.

Science Essay Example SPM

Scientific Essay Example Pdf (Insert

Science Paper Example

Science Project Essay Example

Science Essay Examples for Different Subjects

Science is a vast field that encompasses many different subjects, from biology to physics to chemistry. As a student, you may find yourself tasked with writing a science essay on a subject that you're not particularly familiar with. 

We have provided you with science essay examples for different subjects to help you get started.

Social Science Essay Example

Political Science Essay Example

Environmental Science Essay Example

Health Science Essay Example

Computer Science Essay Example

University Science Essay Examples

Science essays are important part of university-level education. However, different universities may have different requirements and expectations when it comes to writing these essays. 

That's why we've compiled some science essay examples for different universities. You can see what works and what doesn't, and tailor your own writing accordingly.

Scientific Essay Example University

Mcmaster Health Science Essay Example

Cornell Arts And Science Essay Example

Order Essay

Tough Essay Due? Hire Tough Writers!

Structure of a Science Essay

Science essays are a crucial part of many subjects, and learning to structure them effectively is essential for achieving academic success. 

Let’s explore scientific essay structure.

Introduction

The introduction of a science essay should introduce the topic and provide some context for the reader. 

You should explain the purpose of the essay and provide a thesis statement that outlines the main argument you will make in the essay. A good introduction should also capture the reader's interest and motivate them to read on.

Check out these how to start a science essay examples for better understanding:

The advancement of science and technology has transformed the world we live in. From the discovery of electricity to the invention of the internet, our society has made remarkable progress in understanding and utilizing the forces of nature. Science has not only improved our daily lives but also paved the way for groundbreaking innovations and discoveries that have changed the course of history. In this essay, we will explore the significance of science in our modern world and the impact it has on our daily lives. We will also examine the role of science in shaping our future and the ethical considerations that arise from its use. Through this exploration, we hope to gain a deeper understanding of the importance of science and its impact on our world.

Body Paragraphs

The body paragraphs of a science essay should provide evidence to support the thesis statement. You should use scientific evidence, research, and data to support your argument. 

Each paragraph should focus on one key point, and the points should be organized logically to create a coherent argument. It is essential to provide citations for all sources you use in your essay.

Here is an example for you:

One of the most significant impacts of science on our world is the development of new technologies. From smartphones to electric cars, science has led to countless innovations that have made our lives easier and more convenient. However, with these advancements also come ethical considerations. For example, the development of artificial intelligence (AI) has raised concerns about the potential loss of jobs and the ethical implications of relying on machines to make important decisions. Similarly, the use of genetically modified organisms (GMOs) has sparked debates about the safety and environmental impact of altering the genetic makeup of living organisms. As we continue to make scientific advancements, it is essential to consider the ethical implications and ensure that we are using science to benefit society as a whole.

The conclusion of a science essay should summarize the main points of the essay and restate the thesis statement in a compelling manner. 

You should also provide some final thoughts or recommendations based on the evidence presented in the essay. 

The conclusion should be concise and leave a lasting impression on the reader.

In conclusion, science plays a vital role in our modern world. It has led to significant advancements in technology, medicine, and our understanding of the natural world. However, with these advancements come ethical considerations that must be carefully considered. It is essential that we continue to use science to benefit society as a whole and address the challenges facing our world, from climate change to pandemics. Through a greater understanding of science and its impact on our world, we can work towards a brighter future for ourselves and future generations.

Natural Science Essay Topics

There are countless interesting, thought-provoking and problem solving essay topics in science.

Explore some compelling natural science essay topics to inspire your writing.

Science Essay Topics for 5th Graders

  • The importance of recycling for our environment
  • The different types of clouds and how they form
  • How animals hibernate during the winter months
  • The different types of rocks and how they are formed
  • The role of bees in pollination and food production
  • How light travels and how we see objects
  • The properties of magnets and how they work
  • The different stages of stem cell research 
  • The human digestive system and how it works
  • The effects of pollution on our environment and health

Science Essay Topics for 6th Graders

  • The impact of climate change on the planet
  • The different types of energy and how they are produced
  • The importance of water conservation and management
  • The role of artificial intelligence in human life
  • The structure and function of the human respiratory system
  • The properties and uses of acids and bases
  • The effect of light on plant growth and development
  • The differences between renewable and non-renewable energy sources
  • The process of photosynthesis and its importance for life on Earth
  • The impact of technology on the environment and society

Science Essay Topics for 7th Graders

  • The structure and function of the human circulatory system
  • The different types of fossils and how they are formed
  • The impact of natural disasters on the environment and human life
  • The pros and cons of bacteria in our bodies and in the environment
  • The physics of sound and how it travels
  • The effects of air pollution in United States
  • The properties and uses of different types of waves (sound, light, etc.)
  • The process of cell division and its role in growth and repair
  • The structure and function of the human nervous system
  • The different types of ecosystems and their unique characteristics

Paper Due? Why Suffer? That's our Job!

Tips for Writing a Science Essay

Writing a science essay can be challenging, especially if you don't have much experience in writing academic papers. 

However, with the right approach and strategies, you can produce a high-quality science essays. 

Here are some tips to help you write a successful science essay:

Understand the assignment requirements: Before you start writing your essay, make sure you understand the assignment requirements. Read the prompt carefully and make note of any specific guidelines or formatting requirements.

Choose a topic that interests you: Writing about a topic that you find interesting and engaging can make the process enjoyable and rewarding. Consider topics that you have studied in class or that you have a personal interest in.

Conduct thorough research: To write a successful science essay, you need to have a deep understanding of the topic you are writing about. Conduct thorough research using reliable sources such as academic journals, textbooks, and reputable websites.

Develop a clear and concise thesis statement: Your thesis statement should clearly state your argument or position on the topic you are writing about. It should be concise and specific, and should be supported by evidence throughout your essay.

Use evidence to support your claims: When writing a science essay, it's important to use evidence to support your claims and arguments. This can include scientific data, research findings, and expert opinions.

Edit and proofread your essay: Before submitting your essay, make sure to edit and proofread it carefully. Check for spelling and grammatical errors. Ensure that your essay is formatted correctly according to the assignment requirements.

In conclusion, this blog has provided a comprehensive guide to writing a successful science essay. 

By following the tips, students can produce high-quality essays that showcase their understanding of science.

If you're struggling to write a science essay or need additional assistance, CollegeEssay.org is one of the best online essay services to help you out,

Our expert writers have extensive experience in writing science essays for students of all levels. 

So why wait? Contact our science essay writing service today!

Frequently Asked Questions

What are some common mistakes to avoid when writing a science essay.

Some common mistakes to avoid include:

  • Plagiarizing content
  • Using incorrect or unreliable sources
  • Failing to clearly state your thesis
  • Using overly complex language 

How can I make my science essay stand out?

To make your science essay stand out, consider choosing a unique or controversial topic. Using relevant and up-to-date sources, and present your information in a clear and concise manner. You can also consider using visuals such as graphs or charts to enhance your essay.

What should I do if I'm struggling to come up with a topic for my science essay?

If you're struggling to come up with a topic for your science essay, consider discussing potential topics with your instructor or classmates. You can also conduct research online or in academic journals to find inspiration.

How important is research when writing a science essay?

Research is an essential component of writing a science essay. Your essay should be grounded in accurate and reliable scientific information. That is why it's important to conduct thorough research using reputable sources.

Can I use personal anecdotes or experiences in my science essay?

While personal anecdotes or experiences can be engaging, they may not always be relevant to a science essay. It's important to focus on presenting factual information and scientific evidence to support your argument or position.

Caleb S. (Law, Literature)

Caleb S. has extensive experience in writing and holds a Masters from Oxford University. He takes great satisfaction in helping students exceed their academic goals. Caleb always puts the needs of his clients first and is dedicated to providing quality service.

Paper Due? Why Suffer? That’s our Job!

Get Help

Legal & Policies

  • Privacy Policy
  • Cookies Policy
  • Terms of Use
  • Refunds & Cancellations
  • Our Writers
  • Success Stories
  • Our Guarantees
  • Affiliate Program
  • Referral Program
  • AI Essay Writer

Disclaimer: All client orders are completed by our team of highly qualified human writers. The essays and papers provided by us are not to be used for submission but rather as learning models only.

scientific report essay example

Have a language expert improve your writing

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

  • Knowledge Base
  • Research paper

Writing a Research Paper Conclusion | Step-by-Step Guide

Published on October 30, 2022 by Jack Caulfield . Revised on April 13, 2023.

  • Restate the problem statement addressed in the paper
  • Summarize your overall arguments or findings
  • Suggest the key takeaways from your paper

Research paper conclusion

The content of the conclusion varies depending on whether your paper presents the results of original empirical research or constructs an argument through engagement with sources .

Instantly correct all language mistakes in your text

Upload your document to correct all your mistakes in minutes

upload-your-document-ai-proofreader

Table of contents

Step 1: restate the problem, step 2: sum up the paper, step 3: discuss the implications, research paper conclusion examples, frequently asked questions about research paper conclusions.

The first task of your conclusion is to remind the reader of your research problem . You will have discussed this problem in depth throughout the body, but now the point is to zoom back out from the details to the bigger picture.

While you are restating a problem you’ve already introduced, you should avoid phrasing it identically to how it appeared in the introduction . Ideally, you’ll find a novel way to circle back to the problem from the more detailed ideas discussed in the body.

For example, an argumentative paper advocating new measures to reduce the environmental impact of agriculture might restate its problem as follows:

Meanwhile, an empirical paper studying the relationship of Instagram use with body image issues might present its problem like this:

“In conclusion …”

Avoid starting your conclusion with phrases like “In conclusion” or “To conclude,” as this can come across as too obvious and make your writing seem unsophisticated. The content and placement of your conclusion should make its function clear without the need for additional signposting.

Here's why students love Scribbr's proofreading services

Discover proofreading & editing

Having zoomed back in on the problem, it’s time to summarize how the body of the paper went about addressing it, and what conclusions this approach led to.

Depending on the nature of your research paper, this might mean restating your thesis and arguments, or summarizing your overall findings.

Argumentative paper: Restate your thesis and arguments

In an argumentative paper, you will have presented a thesis statement in your introduction, expressing the overall claim your paper argues for. In the conclusion, you should restate the thesis and show how it has been developed through the body of the paper.

Briefly summarize the key arguments made in the body, showing how each of them contributes to proving your thesis. You may also mention any counterarguments you addressed, emphasizing why your thesis holds up against them, particularly if your argument is a controversial one.

Don’t go into the details of your evidence or present new ideas; focus on outlining in broad strokes the argument you have made.

Empirical paper: Summarize your findings

In an empirical paper, this is the time to summarize your key findings. Don’t go into great detail here (you will have presented your in-depth results and discussion already), but do clearly express the answers to the research questions you investigated.

Describe your main findings, even if they weren’t necessarily the ones you expected or hoped for, and explain the overall conclusion they led you to.

Having summed up your key arguments or findings, the conclusion ends by considering the broader implications of your research. This means expressing the key takeaways, practical or theoretical, from your paper—often in the form of a call for action or suggestions for future research.

Argumentative paper: Strong closing statement

An argumentative paper generally ends with a strong closing statement. In the case of a practical argument, make a call for action: What actions do you think should be taken by the people or organizations concerned in response to your argument?

If your topic is more theoretical and unsuitable for a call for action, your closing statement should express the significance of your argument—for example, in proposing a new understanding of a topic or laying the groundwork for future research.

Empirical paper: Future research directions

In a more empirical paper, you can close by either making recommendations for practice (for example, in clinical or policy papers), or suggesting directions for future research.

Whatever the scope of your own research, there will always be room for further investigation of related topics, and you’ll often discover new questions and problems during the research process .

Finish your paper on a forward-looking note by suggesting how you or other researchers might build on this topic in the future and address any limitations of the current paper.

Full examples of research paper conclusions are shown in the tabs below: one for an argumentative paper, the other for an empirical paper.

  • Argumentative paper
  • Empirical paper

While the role of cattle in climate change is by now common knowledge, countries like the Netherlands continually fail to confront this issue with the urgency it deserves. The evidence is clear: To create a truly futureproof agricultural sector, Dutch farmers must be incentivized to transition from livestock farming to sustainable vegetable farming. As well as dramatically lowering emissions, plant-based agriculture, if approached in the right way, can produce more food with less land, providing opportunities for nature regeneration areas that will themselves contribute to climate targets. Although this approach would have economic ramifications, from a long-term perspective, it would represent a significant step towards a more sustainable and resilient national economy. Transitioning to sustainable vegetable farming will make the Netherlands greener and healthier, setting an example for other European governments. Farmers, policymakers, and consumers must focus on the future, not just on their own short-term interests, and work to implement this transition now.

As social media becomes increasingly central to young people’s everyday lives, it is important to understand how different platforms affect their developing self-conception. By testing the effect of daily Instagram use among teenage girls, this study established that highly visual social media does indeed have a significant effect on body image concerns, with a strong correlation between the amount of time spent on the platform and participants’ self-reported dissatisfaction with their appearance. However, the strength of this effect was moderated by pre-test self-esteem ratings: Participants with higher self-esteem were less likely to experience an increase in body image concerns after using Instagram. This suggests that, while Instagram does impact body image, it is also important to consider the wider social and psychological context in which this usage occurs: Teenagers who are already predisposed to self-esteem issues may be at greater risk of experiencing negative effects. Future research into Instagram and other highly visual social media should focus on establishing a clearer picture of how self-esteem and related constructs influence young people’s experiences of these platforms. Furthermore, while this experiment measured Instagram usage in terms of time spent on the platform, observational studies are required to gain more insight into different patterns of usage—to investigate, for instance, whether active posting is associated with different effects than passive consumption of social media content.

If you’re unsure about the conclusion, it can be helpful to ask a friend or fellow student to read your conclusion and summarize the main takeaways.

  • Do they understand from your conclusion what your research was about?
  • Are they able to summarize the implications of your findings?
  • Can they answer your research question based on your conclusion?

You can also get an expert to proofread and feedback your paper with a paper editing service .

Scribbr Citation Checker New

The AI-powered Citation Checker helps you avoid common mistakes such as:

  • Missing commas and periods
  • Incorrect usage of “et al.”
  • Ampersands (&) in narrative citations
  • Missing reference entries

scientific report essay example

The conclusion of a research paper has several key elements you should make sure to include:

  • A restatement of the research problem
  • A summary of your key arguments and/or findings
  • A short discussion of the implications of your research

No, it’s not appropriate to present new arguments or evidence in the conclusion . While you might be tempted to save a striking argument for last, research papers follow a more formal structure than this.

All your findings and arguments should be presented in the body of the text (more specifically in the results and discussion sections if you are following a scientific structure). The conclusion is meant to summarize and reflect on the evidence and arguments you have already presented, not introduce new ones.

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.

Caulfield, J. (2023, April 13). Writing a Research Paper Conclusion | Step-by-Step Guide. Scribbr. Retrieved September 16, 2024, from https://www.scribbr.com/research-paper/research-paper-conclusion/

Is this article helpful?

Jack Caulfield

Jack Caulfield

Other students also liked, writing a research paper introduction | step-by-step guide, how to create a structured research paper outline | example, checklist: writing a great research paper, get unlimited documents corrected.

✔ Free APA citation check included ✔ Unlimited document corrections ✔ Specialized in correcting academic texts

UCI Libraries Mobile Site

  • Langson Library
  • Science Library
  • Grunigen Medical Library
  • Law Library
  • Connect From Off-Campus
  • Accessibility
  • Gateway Study Center

Libaries home page

Email this link

Writing a scientific paper.

  • Writing a lab report

What is a "good" title?

"title checklist" from: how to write a good scientific paper. chris a. mack. spie. 2018., other hints for writing a title.

  • INTRODUCTION
  • LITERATURE CITED
  • Bibliography of guides to scientific writing and presenting
  • Peer Review
  • Presentations
  • Lab Report Writing Guides on the Web

The title will be read by many people. Only a few will read the entire paper, therefore all words in the title should be chosen with care. Too short a title is not helpful to the potential reader. However too long a title can sometimes be even less meaningful. Remember a title is not an abstract. Also a title is not a sentence.

Goals: • Fewest possible words that describe the contents of the paper. • Avoid waste words like "Studies on", or "Investigations on" • Use specific terms rather than general • Watch your word order and syntax • Avoid abbreviations and jargon

 The title should be clear and informative, and should reflect the aim and approach of the work.

 The title should be as specific as possible while still describing the full range of the work. Does the title, seen in isolation, give a full yet concise and specific indication of the work reported?

 Do not mention results or conclusions in the title.

 Avoid: overly clever or punny titles that will not fare well with search engines or international audiences; titles that are too short to be descriptive or too long to be read; jargon, acronyms, or trademarked terms. 

  • Whenever possible, use a declarative rather than a neutral title
  • Don't end your title with a question mark?
  • Begin with the keywords
  • Use verbs instead of abstract nouns
  • Avoid abbrev. in the title

From: How to Write and Illustrate a Scientific Paper (2008)

  • << Previous: Writing a lab report
  • Next: ABSTRACT >>
  • Last Updated: Aug 4, 2023 9:33 AM
  • URL: https://guides.lib.uci.edu/scientificwriting

Off-campus? Please use the Software VPN and choose the group UCIFull to access licensed content. For more information, please Click here

Software VPN is not available for guests, so they may not have access to some content when connecting from off-campus.

  • Science Writing
  • All Applied Health Sciences Databases
  • All Biology Databases
  • All Chemistry Article Databases
  • All Computer Science Databases
  • All Math Databases
  • All Nursing Databases
  • All Physics Databases
  • All Psychology Databases
  • All Sport Management Databases

Profile Photo

Importance of Citation

  • Quoting and Paraphrasing From the University of Wisconsin-Madison's Writer's Handbook.
  • Plagiarism Overview From the OWL (Online Writing Lab) at Purdue.
  • NY Times article: Plagiarism Lines Blur for Students in Digital Age August 1, 2010 Students believe they can use without attribution materials they find on websites where authorship is not stated.

DEI Resources in STEM fields

  • Black Past: STEM people hidden no more

scientific report essay example

What is "Science Writing?"

At first glance, science writing seems simple. It's writing about  topics related to science and can include works of creative nonfiction, nature essays, or scientific articles. But science writing is different from scientific writing. While scientific writing is usually intended for a specialized audience of scientists and researchers, science writing brings important scientific discoveries into the lives of the general public in creative and meaningful ways. Pieces of science writing can be very different from one another and can come in a number of formats. They can be long or short, detailed or generalized. They can be articles, books, videos, essays, podcasts, and more. 

This page serves as a guide for discovering credible science writing sources and connecting various fields of study. Scientists, students, educators, and writers can use it to find articles, videos, books, and other examples of science writing to support their interests. Whether you are an avid reader or an aspiring science writer, this guide is for you!

How to use this guide

  • Use the green side menu to navigate to databases for further scientific research resources. 
  • To read about science topics, explore the tabs in the Science Writing "Reader's Guide."
  • To learn more on writing about science, explore the tabs in Science Writing "Writer's Guide."

Science Writing - Reader's Guide

  • Author Search
  • Periodicals
  • Recent Books and Ebooks
  • Science in the News

Try browsing the library shelves for science writing material in these areas, or explore the Science Writing collection online.

  • Q 130-141 Women in Science
  • Q 143 General Science Biographies 
  • Q 148-149 African American Scientists
  • QA 75-76 Computer Science
  • QB Astronomy 
  • QD 71-142 Analytical Chemistry
  • QD 146-197 Inorganic Chemistry
  • QD 415 -436 Biochemistry
  • QE 701-760 Paleontology 
  • QH Biology/Natural History,
  • QM Human anatomy 
  • QP Physiology
  • QR Microbiology
  • RC 321-576 Psychology/Psychiatry
  • S Agriculture
  • TX Food Science

Searching for science writing materials can be difficult because they span a broad range of topics. Here are some authors to search for to help get you started. 

Nature Writing

  • Susan Fenimore Cooper (1813-1894)
  • Henry David Thoreau (1817-1862)
  • John Muir (1838-1914)

Conservation

  • Aldo Leopold  (1887- 1948)
  • Rachel Carson  (1907-1964)
  • James Lovelock (1919 - )
  • E.O. Wilson (1929 - 2021)
  • Oliver Sacks (1933 - 2015)

Natural Science

  • Carl Sagan (1934 -1996)
  • Richard Dawkins  (1941 - )
  • Stephen Hawking (1942-2018)
  • Michio Kaku (1947- )
  • Dava Sobel  (1947-)
  • Bill Bryson (1951 - )
  • Sy Montgomery (1958 - )
  • Mary Roach (1959 - )
  • Hannah Holmes (1963 - )
  • Peter Wohlleben (1964 - )
  • Carl Zimmer (1966 - ) 
  • Rebecca Skloot (1972 - )
  • Ed Yong (1981 - )
  • Nature This link opens in a new window Journal articles and news in science and technology.
  • Cell Cell is a peer-reviewed scientific journal publishing research papers across a broad range of disciplines within the life sciences including but not limited to cell biology, molecular biology, neuroscience, immunology, virology and microbiology, cancer, human genetics, systems biology, signaling, and disease mechanisms and therapeutics
  • Parkside Journal of Science Parkside journal containing independently conducted research done by students of UW-Parkside, reviews of current scientific articles, interviews of faculty members pertaining to their work.
  • Scientific American Science magazine that covers research, ideas and knowledge in science, health, technology, the environment and society.
  • Science The Science family of journals is published by the American Association for the Advancement of Science (AAAS), the world’s oldest and largest general science organization. Covering a variety of science related topics.
  • Science and Nature Writing Collection Explore more books online in the Science and Nature Writing Collection

Cover Art

  • Access World News Research Collection This link opens in a new window International, national, and regional news. Date coverage varies by newspaper.
  • New York Times Academic Pass This link opens in a new window Anyone with a valid UWP email address can sign up for free access to the NYTimes.com website and smartphone apps. Please read registration instructions. Navigate to the Science page to read the latest science news articles.
  • Points of View Reference Center (EBSCO) This link opens in a new window Sources on contemporary controversial topics. The previous interface includes a list of potential research areas .
  • Science Podcast A weekly podcast featuring the latest scientific research and stories from the journal Science and its sister journals.
  • Nature Podcast The Nature Podcast brings you the best stories from the world of science each week. We cover everything from astronomy to zoology, highlighting the most exciting research from each issue of the Nature journal. We meet the scientists behind the results and provide in-depth analysis from Nature's journalists and editors.
  • Big Biology Big Biology is a podcast that tells the stories of scientists tackling some of the biggest unanswered questions in biology. Hosted by biologists Marty Martin and Art Woods.
  • Body Stuff Podcast created to dispel medical myths and misinformation hosted by Dr. Jen Gunter OB/GYN and pain medicine physician.
  • Hand Made A podcast that explores all things material hosted by materials scientist and engineer, Dr. Anna Ploszajski
  • Hidden Brain Shankar Vedantam uses science and storytelling to reveal the unconscious patterns that drive human behavior, shape our choices and direct our relationships.
  • Introduced A podcast that explores stories of aquatic invasive species in Wisconsin, like rusty crayfish, purple loosestrife and spiny waterfleas.
  • The Pulse The Pulse is a podcast that takes you behind the doors of operating rooms, into the lab with some of the world’s foremost scientists, and back in time to explore life-changing innovations.
  • Radiolab Science journalists explore topics in science from many different angles and viewpoints through in depth interviews and investigative reporting.
  • Science and Nutrition The world’s top scientists explain the latest health, nutrition, and gut health research and translate it into practical advice to improve your health & weight.
  • Science Clear and Vivid Host Alan Alda leads conversations with 10 guests discussing their lives as scientists and their scientific research, each sharing their point of view and experience about basic science and their personal stories – what got them interested in or involved in science and more.
  • Science Friday Science, technology, and other cool stuff. It's brain fun, for curious people
  • Science Vs Science journalists and fact checkers uncover what is fact and what is not.
  • Short Wave Broadcast by NPR that explores discoveries, everyday mysteries, and the science behind the headlines in short episodes every weekday.
  • Story Collider Podcast that explores personal stories about science told by people from all walks of life from scientists and doctors to comedians and artists.
  • Lost Women of Science This podcast uncovers the life and work of women in science who have been forgotten and gives them recognition and credit for their contributions to the scientific world.
  • Deep Look A series that focuses on the very small things of the natural world.
  • Journey to the Microcosmos Real footage of microscopic organisms that inhabit the world narrated by science communicator Hank Green.
  • Kurzgesagt: In a Nut Shell Short animated videos created to explain complex scientific topics such as black holes, neutron stars, or the human immune system.
  • Physics Girl Physics Girl is a YouTube channel created by Dianna Cowern that adventures into the physical sciences with experiments, demonstrations, and cool new discoveries.
  • Science Friday Science, technology, and other cool stuff. It's brain fun, for curious people.
  • Veritasium Videos about science and education explored through interviews with scientists and demonstrations created by Derek Muller.
  • Real Science Short documentaries on life science topics, an offshoot of Real Engineering.

Tutorial

  • Nova Long-running science series that explores issues and discoveries in science and technology.
  • Star Gazers A weekly television series on astronomy.
  • TED Talks Short presentations on topics covering Technology, Entertainment, and Design.

Science Writing - Writer's Guide

  • Education and Professional Organizations
  • Browser Extensions

Cover Art

  • MIT Science Writing Program
  • Madison School of Journalism
  • Alan Alda Center for Communicating Science
  • National Association of Science Writers
  • Open Access Button
  • LibKey Nomad
  • Next: All Applied Health Sciences Databases >>
  • Last Updated: Sep 17, 2024 3:24 PM
  • URL: https://libguides.uwp.edu/sciencewriting

University of Wisconsin-Parkside Library   |    Contact Us 900 Wood Road Kenosha, WI 53141 | (262) 595-3432

Literature and Science: Matthew Arnold’s Analysis

How it works

  • 1 Introduction
  • 2 The Overemphasis on Science and Its Consequences
  • 3 The Unique Contributions of Literature
  • 4 The Vision for an Integrated Educational Approach
  • 5 Conclusion

Introduction

Matthew Arnold, a big name in Victorian poetry and culture, had a lot to say about how literature and science relate. In the 1880s, he gave a lecture called “Literature and Science” that dug into the back-and-forth between these two fields. Arnold’s thoughts came from the time’s cultural and intellectual vibe, a period full of fast scientific progress and a growing love for facts and figures. This essay looks at what Arnold had to say about literature and science, focusing on his idea that literature offers something vital to human growth that science alone can’t provide.

By looking at Arnold’s critique of science’s dominance, his defense of the humanities, and his dream of a balance between literature and science, we can see why his ideas still matter when we talk about balanced education today.

The Overemphasis on Science and Its Consequences

Arnold wasn’t too happy about how science was taking over in his time. He worried that focusing too much on facts might push the humanities to the side. Sure, science and tech are important, but Arnold thought that literature has its own special value. It helps us understand human emotions, ethical issues, and cultural heritage in ways science can’t. Literature, he argued, helps us think critically and reason morally, giving us a fuller picture of the world. Arnold wanted to remind people that society shouldn’t just measure progress by scientific advancements. His critique warned against an education system that only teaches technical skills and forgets about humanistic knowledge.

The Unique Contributions of Literature

Arnold stood up for the humanities by pointing out the unique benefits of literature. He felt that literature, with its universal themes and ability to make us feel empathy, was key to shaping good character and societal values. Literature’s imaginative and reflective nature balances out the analytical side of science. When people read literature, they get a deeper understanding of human experiences and a more nuanced view of life’s problems. Arnold wasn’t just being nostalgic; he truly believed that the humanities are crucial for creating well-rounded people who can think critically and make ethical decisions.

The Vision for an Integrated Educational Approach

Arnold’s idea of blending literature and science in education was pretty forward-thinking. He thought schools should teach both science and the humanities, so students could be both technically skilled and culturally aware. He imagined a world where literature and science work together, making us all smarter and more ethical. This idea fits well with the liberal arts tradition, which aims to develop the whole person, not just one skill set. In today’s world, where education often focuses on specialization, Arnold’s vision is a call to remember the value of a broad education. His ideas are still relevant as we try to prepare students for a world that’s getting more complex and interconnected every day.

Matthew Arnold’s thoughts on literature and science give us a strong case for why the humanities still matter, even in a tech-driven world. His critique of science’s dominance, defense of the humanities, and vision of a balanced education show the need for a mix of technical skills and humanistic understanding. Arnold reminds us that literature, with its ability to dive deep into human experiences and foster critical thinking, is essential for a well-rounded education. As we face the challenges of a fast-changing world, Arnold’s ideas offer valuable guidance. By embracing his call for an integrated educational approach, we can help people navigate modern life’s complexities with both intellectual smarts and ethical sense.

owl

Cite this page

Literature And Science: Matthew Arnold's Analysis. (2024, Sep 17). Retrieved from https://papersowl.com/examples/literature-and-science-matthew-arnolds-analysis/

"Literature And Science: Matthew Arnold's Analysis." PapersOwl.com , 17 Sep 2024, https://papersowl.com/examples/literature-and-science-matthew-arnolds-analysis/

PapersOwl.com. (2024). Literature And Science: Matthew Arnold's Analysis . [Online]. Available at: https://papersowl.com/examples/literature-and-science-matthew-arnolds-analysis/ [Accessed: 18 Sep. 2024]

"Literature And Science: Matthew Arnold's Analysis." PapersOwl.com, Sep 17, 2024. Accessed September 18, 2024. https://papersowl.com/examples/literature-and-science-matthew-arnolds-analysis/

"Literature And Science: Matthew Arnold's Analysis," PapersOwl.com , 17-Sep-2024. [Online]. Available: https://papersowl.com/examples/literature-and-science-matthew-arnolds-analysis/. [Accessed: 18-Sep-2024]

PapersOwl.com. (2024). Literature And Science: Matthew Arnold's Analysis . [Online]. Available at: https://papersowl.com/examples/literature-and-science-matthew-arnolds-analysis/ [Accessed: 18-Sep-2024]

Don't let plagiarism ruin your grade

Hire a writer to get a unique paper crafted to your needs.

owl

Our writers will help you fix any mistakes and get an A+!

Please check your inbox.

You can order an original essay written according to your instructions.

Trusted by over 1 million students worldwide

1. Tell Us Your Requirements

2. Pick your perfect writer

3. Get Your Paper and Pay

Hi! I'm Amy, your personal assistant!

Don't know where to start? Give me your paper requirements and I connect you to an academic expert.

short deadlines

100% Plagiarism-Free

Certified writers

Peer Reviewed

GPT-fabricated scientific papers on Google Scholar: Key features, spread, and implications for preempting evidence manipulation

Article metrics.

CrossRef

CrossRef Citations

Altmetric Score

PDF Downloads

Academic journals, archives, and repositories are seeing an increasing number of questionable research papers clearly produced using generative AI. They are often created with widely available, general-purpose AI applications, most likely ChatGPT, and mimic scientific writing. Google Scholar easily locates and lists these questionable papers alongside reputable, quality-controlled research. Our analysis of a selection of questionable GPT-fabricated scientific papers found in Google Scholar shows that many are about applied, often controversial topics susceptible to disinformation: the environment, health, and computing. The resulting enhanced potential for malicious manipulation of society’s evidence base, particularly in politically divisive domains, is a growing concern.

Swedish School of Library and Information Science, University of Borås, Sweden

Department of Arts and Cultural Sciences, Lund University, Sweden

Division of Environmental Communication, Swedish University of Agricultural Sciences, Sweden

scientific report essay example

Research Questions

  • Where are questionable publications produced with generative pre-trained transformers (GPTs) that can be found via Google Scholar published or deposited?
  • What are the main characteristics of these publications in relation to predominant subject categories?
  • How are these publications spread in the research infrastructure for scholarly communication?
  • How is the role of the scholarly communication infrastructure challenged in maintaining public trust in science and evidence through inappropriate use of generative AI?

research note Summary

  • A sample of scientific papers with signs of GPT-use found on Google Scholar was retrieved, downloaded, and analyzed using a combination of qualitative coding and descriptive statistics. All papers contained at least one of two common phrases returned by conversational agents that use large language models (LLM) like OpenAI’s ChatGPT. Google Search was then used to determine the extent to which copies of questionable, GPT-fabricated papers were available in various repositories, archives, citation databases, and social media platforms.
  • Roughly two-thirds of the retrieved papers were found to have been produced, at least in part, through undisclosed, potentially deceptive use of GPT. The majority (57%) of these questionable papers dealt with policy-relevant subjects (i.e., environment, health, computing), susceptible to influence operations. Most were available in several copies on different domains (e.g., social media, archives, and repositories).
  • Two main risks arise from the increasingly common use of GPT to (mass-)produce fake, scientific publications. First, the abundance of fabricated “studies” seeping into all areas of the research infrastructure threatens to overwhelm the scholarly communication system and jeopardize the integrity of the scientific record. A second risk lies in the increased possibility that convincingly scientific-looking content was in fact deceitfully created with AI tools and is also optimized to be retrieved by publicly available academic search engines, particularly Google Scholar. However small, this possibility and awareness of it risks undermining the basis for trust in scientific knowledge and poses serious societal risks.

Implications

The use of ChatGPT to generate text for academic papers has raised concerns about research integrity. Discussion of this phenomenon is ongoing in editorials, commentaries, opinion pieces, and on social media (Bom, 2023; Stokel-Walker, 2024; Thorp, 2023). There are now several lists of papers suspected of GPT misuse, and new papers are constantly being added. 1 See for example Academ-AI, https://www.academ-ai.info/ , and Retraction Watch, https://retractionwatch.com/papers-and-peer-reviews-with-evidence-of-chatgpt-writing/ . While many legitimate uses of GPT for research and academic writing exist (Huang & Tan, 2023; Kitamura, 2023; Lund et al., 2023), its undeclared use—beyond proofreading—has potentially far-reaching implications for both science and society, but especially for their relationship. It, therefore, seems important to extend the discussion to one of the most accessible and well-known intermediaries between science, but also certain types of misinformation, and the public, namely Google Scholar, also in response to the legitimate concerns that the discussion of generative AI and misinformation needs to be more nuanced and empirically substantiated  (Simon et al., 2023).

Google Scholar, https://scholar.google.com , is an easy-to-use academic search engine. It is available for free, and its index is extensive (Gusenbauer & Haddaway, 2020). It is also often touted as a credible source for academic literature and even recommended in library guides, by media and information literacy initiatives, and fact checkers (Tripodi et al., 2023). However, Google Scholar lacks the transparency and adherence to standards that usually characterize citation databases. Instead, Google Scholar uses automated crawlers, like Google’s web search engine (Martín-Martín et al., 2021), and the inclusion criteria are based on primarily technical standards, allowing any individual author—with or without scientific affiliation—to upload papers to be indexed (Google Scholar Help, n.d.). It has been shown that Google Scholar is susceptible to manipulation through citation exploits (Antkare, 2020) and by providing access to fake scientific papers (Dadkhah et al., 2017). A large part of Google Scholar’s index consists of publications from established scientific journals or other forms of quality-controlled, scholarly literature. However, the index also contains a large amount of gray literature, including student papers, working papers, reports, preprint servers, and academic networking sites, as well as material from so-called “questionable” academic journals, including paper mills. The search interface does not offer the possibility to filter the results meaningfully by material type, publication status, or form of quality control, such as limiting the search to peer-reviewed material.

To understand the occurrence of ChatGPT (co-)authored work in Google Scholar’s index, we scraped it for publications, including one of two common ChatGPT responses (see Appendix A) that we encountered on social media and in media reports (DeGeurin, 2024). The results of our descriptive statistical analyses showed that around 62% did not declare the use of GPTs. Most of these GPT-fabricated papers were found in non-indexed journals and working papers, but some cases included research published in mainstream scientific journals and conference proceedings. 2 Indexed journals mean scholarly journals indexed by abstract and citation databases such as Scopus and Web of Science, where the indexation implies journals with high scientific quality. Non-indexed journals are journals that fall outside of this indexation. More than half (57%) of these GPT-fabricated papers concerned policy-relevant subject areas susceptible to influence operations. To avoid increasing the visibility of these publications, we abstained from referencing them in this research note. However, we have made the data available in the Harvard Dataverse repository.

The publications were related to three issue areas—health (14.5%), environment (19.5%) and computing (23%)—with key terms such “healthcare,” “COVID-19,” or “infection”for health-related papers, and “analysis,” “sustainable,” and “global” for environment-related papers. In several cases, the papers had titles that strung together general keywords and buzzwords, thus alluding to very broad and current research. These terms included “biology,” “telehealth,” “climate policy,” “diversity,” and “disrupting,” to name just a few.  While the study’s scope and design did not include a detailed analysis of which parts of the articles included fabricated text, our dataset did contain the surrounding sentences for each occurrence of the suspicious phrases that formed the basis for our search and subsequent selection. Based on that, we can say that the phrases occurred in most sections typically found in scientific publications, including the literature review, methods, conceptual and theoretical frameworks, background, motivation or societal relevance, and even discussion. This was confirmed during the joint coding, where we read and discussed all articles. It became clear that not just the text related to the telltale phrases was created by GPT, but that almost all articles in our sample of questionable articles likely contained traces of GPT-fabricated text everywhere.

Evidence hacking and backfiring effects

Generative pre-trained transformers (GPTs) can be used to produce texts that mimic scientific writing. These texts, when made available online—as we demonstrate—leak into the databases of academic search engines and other parts of the research infrastructure for scholarly communication. This development exacerbates problems that were already present with less sophisticated text generators (Antkare, 2020; Cabanac & Labbé, 2021). Yet, the public release of ChatGPT in 2022, together with the way Google Scholar works, has increased the likelihood of lay people (e.g., media, politicians, patients, students) coming across questionable (or even entirely GPT-fabricated) papers and other problematic research findings. Previous research has emphasized that the ability to determine the value and status of scientific publications for lay people is at stake when misleading articles are passed off as reputable (Haider & Åström, 2017) and that systematic literature reviews risk being compromised (Dadkhah et al., 2017). It has also been highlighted that Google Scholar, in particular, can be and has been exploited for manipulating the evidence base for politically charged issues and to fuel conspiracy narratives (Tripodi et al., 2023). Both concerns are likely to be magnified in the future, increasing the risk of what we suggest calling evidence hacking —the strategic and coordinated malicious manipulation of society’s evidence base.

The authority of quality-controlled research as evidence to support legislation, policy, politics, and other forms of decision-making is undermined by the presence of undeclared GPT-fabricated content in publications professing to be scientific. Due to the large number of archives, repositories, mirror sites, and shadow libraries to which they spread, there is a clear risk that GPT-fabricated, questionable papers will reach audiences even after a possible retraction. There are considerable technical difficulties involved in identifying and tracing computer-fabricated papers (Cabanac & Labbé, 2021; Dadkhah et al., 2023; Jones, 2024), not to mention preventing and curbing their spread and uptake.

However, as the rise of the so-called anti-vaxx movement during the COVID-19 pandemic and the ongoing obstruction and denial of climate change show, retracting erroneous publications often fuels conspiracies and increases the following of these movements rather than stopping them. To illustrate this mechanism, climate deniers frequently question established scientific consensus by pointing to other, supposedly scientific, studies that support their claims. Usually, these are poorly executed, not peer-reviewed, based on obsolete data, or even fraudulent (Dunlap & Brulle, 2020). A similar strategy is successful in the alternative epistemic world of the global anti-vaccination movement (Carrion, 2018) and the persistence of flawed and questionable publications in the scientific record already poses significant problems for health research, policy, and lawmakers, and thus for society as a whole (Littell et al., 2024). Considering that a person’s support for “doing your own research” is associated with increased mistrust in scientific institutions (Chinn & Hasell, 2023), it will be of utmost importance to anticipate and consider such backfiring effects already when designing a technical solution, when suggesting industry or legal regulation, and in the planning of educational measures.

Recommendations

Solutions should be based on simultaneous considerations of technical, educational, and regulatory approaches, as well as incentives, including social ones, across the entire research infrastructure. Paying attention to how these approaches and incentives relate to each other can help identify points and mechanisms for disruption. Recognizing fraudulent academic papers must happen alongside understanding how they reach their audiences and what reasons there might be for some of these papers successfully “sticking around.” A possible way to mitigate some of the risks associated with GPT-fabricated scholarly texts finding their way into academic search engine results would be to provide filtering options for facets such as indexed journals, gray literature, peer-review, and similar on the interface of publicly available academic search engines. Furthermore, evaluation tools for indexed journals 3 Such as LiU Journal CheckUp, https://ep.liu.se/JournalCheckup/default.aspx?lang=eng . could be integrated into the graphical user interfaces and the crawlers of these academic search engines. To enable accountability, it is important that the index (database) of such a search engine is populated according to criteria that are transparent, open to scrutiny, and appropriate to the workings of  science and other forms of academic research. Moreover, considering that Google Scholar has no real competitor, there is a strong case for establishing a freely accessible, non-specialized academic search engine that is not run for commercial reasons but for reasons of public interest. Such measures, together with educational initiatives aimed particularly at policymakers, science communicators, journalists, and other media workers, will be crucial to reducing the possibilities for and effects of malicious manipulation or evidence hacking. It is important not to present this as a technical problem that exists only because of AI text generators but to relate it to the wider concerns in which it is embedded. These range from a largely dysfunctional scholarly publishing system (Haider & Åström, 2017) and academia’s “publish or perish” paradigm to Google’s near-monopoly and ideological battles over the control of information and ultimately knowledge. Any intervention is likely to have systemic effects; these effects need to be considered and assessed in advance and, ideally, followed up on.

Our study focused on a selection of papers that were easily recognizable as fraudulent. We used this relatively small sample as a magnifying glass to examine, delineate, and understand a problem that goes beyond the scope of the sample itself, which however points towards larger concerns that require further investigation. The work of ongoing whistleblowing initiatives 4 Such as Academ-AI, https://www.academ-ai.info/ , and Retraction Watch, https://retractionwatch.com/papers-and-peer-reviews-with-evidence-of-chatgpt-writing/ . , recent media reports of journal closures (Subbaraman, 2024), or GPT-related changes in word use and writing style (Cabanac et al., 2021; Stokel-Walker, 2024) suggest that we only see the tip of the iceberg. There are already more sophisticated cases (Dadkhah et al., 2023) as well as cases involving fabricated images (Gu et al., 2022). Our analysis shows that questionable and potentially manipulative GPT-fabricated papers permeate the research infrastructure and are likely to become a widespread phenomenon. Our findings underline that the risk of fake scientific papers being used to maliciously manipulate evidence (see Dadkhah et al., 2017) must be taken seriously. Manipulation may involve undeclared automatic summaries of texts, inclusion in literature reviews, explicit scientific claims, or the concealment of errors in studies so that they are difficult to detect in peer review. However, the mere possibility of these things happening is a significant risk in its own right that can be strategically exploited and will have ramifications for trust in and perception of science. Society’s methods of evaluating sources and the foundations of media and information literacy are under threat and public trust in science is at risk of further erosion, with far-reaching consequences for society in dealing with information disorders. To address this multifaceted problem, we first need to understand why it exists and proliferates.

Finding 1: 139 GPT-fabricated, questionable papers were found and listed as regular results on the Google Scholar results page. Non-indexed journals dominate.

Most questionable papers we found were in non-indexed journals or were working papers, but we did also find some in established journals, publications, conferences, and repositories. We found a total of 139 papers with a suspected deceptive use of ChatGPT or similar LLM applications (see Table 1). Out of these, 19 were in indexed journals, 89 were in non-indexed journals, 19 were student papers found in university databases, and 12 were working papers (mostly in preprint databases). Table 1 divides these papers into categories. Health and environment papers made up around 34% (47) of the sample. Of these, 66% were present in non-indexed journals.

Indexed journals*534719
Non-indexed journals1818134089
Student papers4311119
Working papers532212
Total32272060139

Finding 2: GPT-fabricated, questionable papers are disseminated online, permeating the research infrastructure for scholarly communication, often in multiple copies. Applied topics with practical implications dominate.

The 20 papers concerning health-related issues are distributed across 20 unique domains, accounting for 46 URLs. The 27 papers dealing with environmental issues can be found across 26 unique domains, accounting for 56 URLs.  Most of the identified papers exist in multiple copies and have already spread to several archives, repositories, and social media. It would be difficult, or impossible, to remove them from the scientific record.

As apparent from Table 2, GPT-fabricated, questionable papers are seeping into most parts of the online research infrastructure for scholarly communication. Platforms on which identified papers have appeared include ResearchGate, ORCiD, Journal of Population Therapeutics and Clinical Pharmacology (JPTCP), Easychair, Frontiers, the Institute of Electrical and Electronics Engineer (IEEE), and X/Twitter. Thus, even if they are retracted from their original source, it will prove very difficult to track, remove, or even just mark them up on other platforms. Moreover, unless regulated, Google Scholar will enable their continued and most likely unlabeled discoverability.

Environmentresearchgate.net (13)orcid.org (4)easychair.org (3)ijope.com* (3)publikasiindonesia.id (3)
Healthresearchgate.net (15)ieee.org (4)twitter.com (3)jptcp.com** (2)frontiersin.org
(2)

A word rain visualization (Centre for Digital Humanities Uppsala, 2023), which combines word prominences through TF-IDF 5 Term frequency–inverse document frequency , a method for measuring the significance of a word in a document compared to its frequency across all documents in a collection. scores with semantic similarity of the full texts of our sample of GPT-generated articles that fall into the “Environment” and “Health” categories, reflects the two categories in question. However, as can be seen in Figure 1, it also reveals overlap and sub-areas. The y-axis shows word prominences through word positions and font sizes, while the x-axis indicates semantic similarity. In addition to a certain amount of overlap, this reveals sub-areas, which are best described as two distinct events within the word rain. The event on the left bundles terms related to the development and management of health and healthcare with “challenges,” “impact,” and “potential of artificial intelligence”emerging as semantically related terms. Terms related to research infrastructures, environmental, epistemic, and technological concepts are arranged further down in the same event (e.g., “system,” “climate,” “understanding,” “knowledge,” “learning,” “education,” “sustainable”). A second distinct event further to the right bundles terms associated with fish farming and aquatic medicinal plants, highlighting the presence of an aquaculture cluster.  Here, the prominence of groups of terms such as “used,” “model,” “-based,” and “traditional” suggests the presence of applied research on these topics. The two events making up the word rain visualization, are linked by a less dominant but overlapping cluster of terms related to “energy” and “water.”

scientific report essay example

The bar chart of the terms in the paper subset (see Figure 2) complements the word rain visualization by depicting the most prominent terms in the full texts along the y-axis. Here, word prominences across health and environment papers are arranged descendingly, where values outside parentheses are TF-IDF values (relative frequencies) and values inside parentheses are raw term frequencies (absolute frequencies).

scientific report essay example

Finding 3: Google Scholar presents results from quality-controlled and non-controlled citation databases on the same interface, providing unfiltered access to GPT-fabricated questionable papers.

Google Scholar’s central position in the publicly accessible scholarly communication infrastructure, as well as its lack of standards, transparency, and accountability in terms of inclusion criteria, has potentially serious implications for public trust in science. This is likely to exacerbate the already-known potential to exploit Google Scholar for evidence hacking (Tripodi et al., 2023) and will have implications for any attempts to retract or remove fraudulent papers from their original publication venues. Any solution must consider the entirety of the research infrastructure for scholarly communication and the interplay of different actors, interests, and incentives.

We searched and scraped Google Scholar using the Python library Scholarly (Cholewiak et al., 2023) for papers that included specific phrases known to be common responses from ChatGPT and similar applications with the same underlying model (GPT3.5 or GPT4): “as of my last knowledge update” and/or “I don’t have access to real-time data” (see Appendix A). This facilitated the identification of papers that likely used generative AI to produce text, resulting in 227 retrieved papers. The papers’ bibliographic information was automatically added to a spreadsheet and downloaded into Zotero. 6 An open-source reference manager, https://zotero.org .

We employed multiple coding (Barbour, 2001) to classify the papers based on their content. First, we jointly assessed whether the paper was suspected of fraudulent use of ChatGPT (or similar) based on how the text was integrated into the papers and whether the paper was presented as original research output or the AI tool’s role was acknowledged. Second, in analyzing the content of the papers, we continued the multiple coding by classifying the fraudulent papers into four categories identified during an initial round of analysis—health, environment, computing, and others—and then determining which subjects were most affected by this issue (see Table 1). Out of the 227 retrieved papers, 88 papers were written with legitimate and/or declared use of GPTs (i.e., false positives, which were excluded from further analysis), and 139 papers were written with undeclared and/or fraudulent use (i.e., true positives, which were included in further analysis). The multiple coding was conducted jointly by all authors of the present article, who collaboratively coded and cross-checked each other’s interpretation of the data simultaneously in a shared spreadsheet file. This was done to single out coding discrepancies and settle coding disagreements, which in turn ensured methodological thoroughness and analytical consensus (see Barbour, 2001). Redoing the category coding later based on our established coding schedule, we achieved an intercoder reliability (Cohen’s kappa) of 0.806 after eradicating obvious differences.

The ranking algorithm of Google Scholar prioritizes highly cited and older publications (Martín-Martín et al., 2016). Therefore, the position of the articles on the search engine results pages was not particularly informative, considering the relatively small number of results in combination with the recency of the publications. Only the query “as of my last knowledge update” had more than two search engine result pages. On those, questionable articles with undeclared use of GPTs were evenly distributed across all result pages (min: 4, max: 9, mode: 8), with the proportion of undeclared use being slightly higher on average on later search result pages.

To understand how the papers making fraudulent use of generative AI were disseminated online, we programmatically searched for the paper titles (with exact string matching) in Google Search from our local IP address (see Appendix B) using the googlesearch – python library(Vikramaditya, 2020). We manually verified each search result to filter out false positives—results that were not related to the paper—and then compiled the most prominent URLs by field. This enabled the identification of other platforms through which the papers had been spread. We did not, however, investigate whether copies had spread into SciHub or other shadow libraries, or if they were referenced in Wikipedia.

We used descriptive statistics to count the prevalence of the number of GPT-fabricated papers across topics and venues and top domains by subject. The pandas software library for the Python programming language (The pandas development team, 2024) was used for this part of the analysis. Based on the multiple coding, paper occurrences were counted in relation to their categories, divided into indexed journals, non-indexed journals, student papers, and working papers. The schemes, subdomains, and subdirectories of the URL strings were filtered out while top-level domains and second-level domains were kept, which led to normalizing domain names. This, in turn, allowed the counting of domain frequencies in the environment and health categories. To distinguish word prominences and meanings in the environment and health-related GPT-fabricated questionable papers, a semantically-aware word cloud visualization was produced through the use of a word rain (Centre for Digital Humanities Uppsala, 2023) for full-text versions of the papers. Font size and y-axis positions indicate word prominences through TF-IDF scores for the environment and health papers (also visualized in a separate bar chart with raw term frequencies in parentheses), and words are positioned along the x-axis to reflect semantic similarity (Skeppstedt et al., 2024), with an English Word2vec skip gram model space (Fares et al., 2017). An English stop word list was used, along with a manually produced list including terms such as “https,” “volume,” or “years.”

  • Artificial Intelligence
  • / Search engines

Cite this Essay

Haider, J., Söderström, K. R., Ekström, B., & Rödl, M. (2024). GPT-fabricated scientific papers on Google Scholar: Key features, spread, and implications for preempting evidence manipulation. Harvard Kennedy School (HKS) Misinformation Review . https://doi.org/10.37016/mr-2020-156

  • / Appendix B

Bibliography

Antkare, I. (2020). Ike Antkare, his publications, and those of his disciples. In M. Biagioli & A. Lippman (Eds.), Gaming the metrics (pp. 177–200). The MIT Press. https://doi.org/10.7551/mitpress/11087.003.0018

Barbour, R. S. (2001). Checklists for improving rigour in qualitative research: A case of the tail wagging the dog? BMJ , 322 (7294), 1115–1117. https://doi.org/10.1136/bmj.322.7294.1115

Bom, H.-S. H. (2023). Exploring the opportunities and challenges of ChatGPT in academic writing: A roundtable discussion. Nuclear Medicine and Molecular Imaging , 57 (4), 165–167. https://doi.org/10.1007/s13139-023-00809-2

Cabanac, G., & Labbé, C. (2021). Prevalence of nonsensical algorithmically generated papers in the scientific literature. Journal of the Association for Information Science and Technology , 72 (12), 1461–1476. https://doi.org/10.1002/asi.24495

Cabanac, G., Labbé, C., & Magazinov, A. (2021). Tortured phrases: A dubious writing style emerging in science. Evidence of critical issues affecting established journals . arXiv. https://doi.org/10.48550/arXiv.2107.06751

Carrion, M. L. (2018). “You need to do your research”: Vaccines, contestable science, and maternal epistemology. Public Understanding of Science , 27 (3), 310–324. https://doi.org/10.1177/0963662517728024

Centre for Digital Humanities Uppsala (2023). CDHUppsala/word-rain [Computer software]. https://github.com/CDHUppsala/word-rain

Chinn, S., & Hasell, A. (2023). Support for “doing your own research” is associated with COVID-19 misperceptions and scientific mistrust. Harvard Kennedy School (HSK) Misinformation Review, 4 (3). https://doi.org/10.37016/mr-2020-117

Cholewiak, S. A., Ipeirotis, P., Silva, V., & Kannawadi, A. (2023). SCHOLARLY: Simple access to Google Scholar authors and citation using Python (1.5.0) [Computer software]. https://doi.org/10.5281/zenodo.5764801

Dadkhah, M., Lagzian, M., & Borchardt, G. (2017). Questionable papers in citation databases as an issue for literature review. Journal of Cell Communication and Signaling , 11 (2), 181–185. https://doi.org/10.1007/s12079-016-0370-6

Dadkhah, M., Oermann, M. H., Hegedüs, M., Raman, R., & Dávid, L. D. (2023). Detection of fake papers in the era of artificial intelligence. Diagnosis , 10 (4), 390–397. https://doi.org/10.1515/dx-2023-0090

DeGeurin, M. (2024, March 19). AI-generated nonsense is leaking into scientific journals. Popular Science. https://www.popsci.com/technology/ai-generated-text-scientific-journals/

Dunlap, R. E., & Brulle, R. J. (2020). Sources and amplifiers of climate change denial. In D.C. Holmes & L. M. Richardson (Eds.), Research handbook on communicating climate change (pp. 49–61). Edward Elgar Publishing. https://doi.org/10.4337/9781789900408.00013

Fares, M., Kutuzov, A., Oepen, S., & Velldal, E. (2017). Word vectors, reuse, and replicability: Towards a community repository of large-text resources. In J. Tiedemann & N. Tahmasebi (Eds.), Proceedings of the 21st Nordic Conference on Computational Linguistics (pp. 271–276). Association for Computational Linguistics. https://aclanthology.org/W17-0237

Google Scholar Help. (n.d.). Inclusion guidelines for webmasters . https://scholar.google.com/intl/en/scholar/inclusion.html

Gu, J., Wang, X., Li, C., Zhao, J., Fu, W., Liang, G., & Qiu, J. (2022). AI-enabled image fraud in scientific publications. Patterns , 3 (7), 100511. https://doi.org/10.1016/j.patter.2022.100511

Gusenbauer, M., & Haddaway, N. R. (2020). Which academic search systems are suitable for systematic reviews or meta-analyses? Evaluating retrieval qualities of Google Scholar, PubMed, and 26 other resources. Research Synthesis Methods , 11 (2), 181–217.   https://doi.org/10.1002/jrsm.1378

Haider, J., & Åström, F. (2017). Dimensions of trust in scholarly communication: Problematizing peer review in the aftermath of John Bohannon’s “Sting” in science. Journal of the Association for Information Science and Technology , 68 (2), 450–467. https://doi.org/10.1002/asi.23669

Huang, J., & Tan, M. (2023). The role of ChatGPT in scientific communication: Writing better scientific review articles. American Journal of Cancer Research , 13 (4), 1148–1154. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10164801/

Jones, N. (2024). How journals are fighting back against a wave of questionable images. Nature , 626 (8000), 697–698. https://doi.org/10.1038/d41586-024-00372-6

Kitamura, F. C. (2023). ChatGPT is shaping the future of medical writing but still requires human judgment. Radiology , 307 (2), e230171. https://doi.org/10.1148/radiol.230171

Littell, J. H., Abel, K. M., Biggs, M. A., Blum, R. W., Foster, D. G., Haddad, L. B., Major, B., Munk-Olsen, T., Polis, C. B., Robinson, G. E., Rocca, C. H., Russo, N. F., Steinberg, J. R., Stewart, D. E., Stotland, N. L., Upadhyay, U. D., & Ditzhuijzen, J. van. (2024). Correcting the scientific record on abortion and mental health outcomes. BMJ , 384 , e076518. https://doi.org/10.1136/bmj-2023-076518

Lund, B. D., Wang, T., Mannuru, N. R., Nie, B., Shimray, S., & Wang, Z. (2023). ChatGPT and a new academic reality: Artificial Intelligence-written research papers and the ethics of the large language models in scholarly publishing. Journal of the Association for Information Science and Technology, 74 (5), 570–581. https://doi.org/10.1002/asi.24750

Martín-Martín, A., Orduna-Malea, E., Ayllón, J. M., & Delgado López-Cózar, E. (2016). Back to the past: On the shoulders of an academic search engine giant. Scientometrics , 107 , 1477–1487. https://doi.org/10.1007/s11192-016-1917-2

Martín-Martín, A., Thelwall, M., Orduna-Malea, E., & Delgado López-Cózar, E. (2021). Google Scholar, Microsoft Academic, Scopus, Dimensions, Web of Science, and OpenCitations’ COCI: A multidisciplinary comparison of coverage via citations. Scientometrics , 126 (1), 871–906. https://doi.org/10.1007/s11192-020-03690-4

Simon, F. M., Altay, S., & Mercier, H. (2023). Misinformation reloaded? Fears about the impact of generative AI on misinformation are overblown. Harvard Kennedy School (HKS) Misinformation Review, 4 (5). https://doi.org/10.37016/mr-2020-127

Skeppstedt, M., Ahltorp, M., Kucher, K., & Lindström, M. (2024). From word clouds to Word Rain: Revisiting the classic word cloud to visualize climate change texts. Information Visualization , 23 (3), 217–238. https://doi.org/10.1177/14738716241236188

Swedish Research Council. (2017). Good research practice. Vetenskapsrådet.

Stokel-Walker, C. (2024, May 1.). AI Chatbots Have Thoroughly Infiltrated Scientific Publishing . Scientific American. https://www.scientificamerican.com/article/chatbots-have-thoroughly-infiltrated-scientific-publishing/

Subbaraman, N. (2024, May 14). Flood of fake science forces multiple journal closures: Wiley to shutter 19 more journals, some tainted by fraud. The Wall Street Journal . https://www.wsj.com/science/academic-studies-research-paper-mills-journals-publishing-f5a3d4bc

The pandas development team. (2024). pandas-dev/pandas: Pandas (v2.2.2) [Computer software]. Zenodo. https://doi.org/10.5281/zenodo.10957263

Thorp, H. H. (2023). ChatGPT is fun, but not an author. Science , 379 (6630), 313–313. https://doi.org/10.1126/science.adg7879

Tripodi, F. B., Garcia, L. C., & Marwick, A. E. (2023). ‘Do your own research’: Affordance activation and disinformation spread. Information, Communication & Society , 27 (6), 1212–1228. https://doi.org/10.1080/1369118X.2023.2245869

Vikramaditya, N. (2020). Nv7-GitHub/googlesearch [Computer software]. https://github.com/Nv7-GitHub/googlesearch

This research has been supported by Mistra, the Swedish Foundation for Strategic Environmental Research, through the research program Mistra Environmental Communication (Haider, Ekström, Rödl) and the Marcus and Amalia Wallenberg Foundation [2020.0004] (Söderström).

Competing Interests

The authors declare no competing interests.

The research described in this article was carried out under Swedish legislation. According to the relevant EU and Swedish legislation (2003:460) on the ethical review of research involving humans (“Ethical Review Act”), the research reported on here is not subject to authorization by the Swedish Ethical Review Authority (“etikprövningsmyndigheten”) (SRC, 2017).

This is an open access article distributed under the terms of the Creative Commons Attribution License , which permits unrestricted use, distribution, and reproduction in any medium, provided that the original author and source are properly credited.

Data Availability

All data needed to replicate this study are available at the Harvard Dataverse: https://doi.org/10.7910/DVN/WUVD8X

Acknowledgements

The authors wish to thank two anonymous reviewers for their valuable comments on the article manuscript as well as the editorial group of Harvard Kennedy School (HKS) Misinformation Review for their thoughtful feedback and input.

Search within the TIB website or find specialist literature and information in the TIB Portal.

The TIB Portal allows you to search the library's own holdings and other data sources simultaneously. By restricting the search to the TIB catalogue, you can search exclusively for printed and digital publications in the entire stock of the TIB library.

Book Review Article (English)

  • ISSN: 1035-8811 , 1757-6547
  • Article (Journal)  /  Electronic Resource

How to get this title?

Show citation formats, export, share and cite, pricing information, please choose your delivery country and your customer group.

*   Mandatory field

More details on this result

  • Title: Book Review Article
  • Published in: The Australian Journal of Anthropology ; 11, 1 ; 59-77
  • Publisher: Blackwell Publishing Ltd
  • New search for: Blackwell Publishing Ltd
  • Publication date: 2000-04-01
  • Size: 19 pages
  • DOI: https://doi.org/10.1111/j.1835-9310.2000.tb00263.x
  • Type of media: Article (Journal)
  • Type of material: Electronic Resource
  • Language: English
  • Source: Wiley

Table of contents

Table of contents – volume 11, issue 1.

Show all volumes and issues

The tables of contents are generated automatically and are based on the data records of the individual contributions available in the index of the TIB portal. The display of the Tables of Contents may therefore be incomplete.

Similar titles

IMAGES

  1. 43+ Research Paper Examples

    scientific report essay example

  2. Writing A Science Essay

    scientific report essay example

  3. How to Write a Research Paper Outline With Examples?

    scientific report essay example

  4. FREE 8+ Sample Scientific Reports in PDF

    scientific report essay example

  5. How to Write a Scientific Paper

    scientific report essay example

  6. 40 Lab Report Templates & Format Examples

    scientific report essay example

VIDEO

  1. How to write Scientific essay for an undergraduate

  2. How to Write an MBA Dissertation ?

  3. How to Write a Scientific Research Paper

  4. how to write a scientific report

  5. Introduction to report writing using the Scientific Method ENG

  6. Writing a Scientific Report

COMMENTS

  1. How to Write a Scientific Report

    A scientific report documents all aspects of an experimental investigation. This includes: A title; The aim of the experiment; The hypothesis; An introduction to the relevant background theory; The methods used; The results; A discussion of the results; The conclusion; Scientific reports allow their readers to understand the experiment without ...

  2. Scientific Reports

    What this handout is about. This handout provides a general guide to writing reports about scientific research you've performed. In addition to describing the conventional rules about the format and content of a lab report, we'll also attempt to convey why these rules exist, so you'll get a clearer, more dependable idea of how to approach this writing situation.

  3. Scientific Writing Made Easy: A Step‐by‐Step Guide to Undergraduate

    The guide breaks down the scientific writing process into easily digestible pieces, providing concrete examples that students can refer to when preparing a scientific manuscript or laboratory report. By increasing undergraduate exposure to the scientific writing process, we hope to better prepare undergraduates for graduate school and ...

  4. How To Write A Lab Report

    Introduction. Your lab report introduction should set the scene for your experiment. One way to write your introduction is with a funnel (an inverted triangle) structure: Start with the broad, general research topic. Narrow your topic down your specific study focus. End with a clear research question.

  5. PDF Tutorial Essays for Science Subjects

    Tutorial Essays for Science Subjects. This guide is designed to provide help and advice on scientific writing. Although students studying Medical and Life Sciences are most likely to have to write essays for tutorials at Oxford, it is important all scientists learn to write clearly and concisely to present their data and conclusions.

  6. Sample Papers

    The following two sample papers were published in annotated form in the Publication Manual and are reproduced here as PDFs for your ease of use. The annotations draw attention to content and formatting and provide the relevant sections of the Publication Manual (7th ed.) to consult for more information.. Student sample paper with annotations (PDF, 5MB)

  7. PDF WRITING A SCIENTIFIC ESSAY

    To give an example For example, for instance, an example of this is, a further instance of this is, To list ideas in order of time First, first of all, first and foremost, second, more important, most important, more significantly, above all, most of all, concurrently, an additional To introduce an explanation or make a stronger statement

  8. The Writing Center

    This format is often used for lab reports as well as for reporting any planned, systematic research in the social sciences, natural sciences, or engineering and computer sciences. Introduction - Make a case for your research. The introduction explains why this research is important or necessary or important. Begin by describing the problem or ...

  9. How to Write a Scientific Paper: Practical Guidelines

    As an example, the number of reports and reviews on obesity and diabetes has increased from 400 to close to 4000/year and 50 to 600/year respectively over a period of 20 years (Figure 3). The present article, essentially based on TA Lang's guide for writing a scientific paper [ 1 ], will summarize the steps involved in the process of writing ...

  10. PDF The Complete Guide to Writing a Report for a Scientific ...

    Generally, a report for a lab experiment comprises of a few essential sections that are common to all. However, depending on the type of experiment or the methodology used, there could be variations in the basic structure. Title Like any other formal document, the lab report should begin with a concise but insightful title for the experiment.

  11. Writing a scientific report

    In this video, a language and learning adviser provides some useful language tips for writing a scientific paper. In summary, these tips are: Be clear about the purpose of the paper. Use precise language. Be aware of your use of verb tense (past tense is often used, as you are reporting on past events in the lab/field).

  12. Writing an Introduction for a Scientific Paper

    Dr. Michelle Harris, Dr. Janet Batzli,Biocore. This section provides guidelines on how to construct a solid introduction to a scientific paper including background information, study question, biological rationale, hypothesis, and general approach. If the Introduction is done well, there should be no question in the reader's mind why and on ...

  13. Research Guides: Writing a Scientific Paper: RESULTS

    Chris A. Mack. SPIE. 2018. Present the results of the paper, in logical order, using tables and graphs as necessary. Explain the results and show how they help to answer the research questions posed in the Introduction. Evidence does not explain itself; the results must be presented and then explained. Avoid: presenting results that are never ...

  14. PDF How to Write Paper in Scientific Journal Style and Format

    Lewiston, ME. v. 10‐2014. This is a reference sheet to help you remember the common format we expect you to use on your formal lab write‐ups. Refer to the "How to Write Guide" for the details. Other than the title, use 12 point type, preferably Calibri, Times New Roman, or Courier.

  15. Scientific Lab Reports

    Writing a Lab Report. Writing a scientific lab report is significantly different from writing for other classes like philosophy, English, and history. The most prominent form of writing in biology, chemistry, and environmental science is the lab report, which is a formally written description of results and discoveries found in an experiment.

  16. How to Write a Report: A Guide to Report Formats with Examples

    1 Choose a topic based on the assignment. Before you start writing, you need to pick the topic of your report. Often, the topic is assigned for you, as with most business reports, or predetermined by the nature of your work, as with scientific reports. If that's the case, you can ignore this step and move on.

  17. How to Write Discussions and Conclusions

    Begin with a clear statement of the principal findings. This will reinforce the main take-away for the reader and set up the rest of the discussion. Explain why the outcomes of your study are important to the reader. Discuss the implications of your findings realistically based on previous literature, highlighting both the strengths and ...

  18. Top 15 Science Essay Examples for Students

    The properties and uses of acids and bases. The effect of light on plant growth and development. The differences between renewable and non-renewable energy sources. The process of photosynthesis and its importance for life on Earth. The impact of technology on the environment and society.

  19. Writing a Research Paper Conclusion

    Table of contents. Step 1: Restate the problem. Step 2: Sum up the paper. Step 3: Discuss the implications. Research paper conclusion examples. Frequently asked questions about research paper conclusions.

  20. Research Guides: Writing a Scientific Paper: TITLE

    However too long a title can sometimes be even less meaningful. Remember a title is not an abstract. Also a title is not a sentence. Goals: • Fewest possible words that describe the contents of the paper. • Avoid waste words like "Studies on", or "Investigations on". • Use specific terms rather than general. • Watch your word order and ...

  21. Science Writing

    What is "Science Writing?" At first glance, science writing seems simple. It's writing about topics related to science and can include works of creative nonfiction, nature essays, or scientific articles. But science writing is different from scientific writing. While scientific writing is usually intended for a specialized audience of scientists and researchers, science writing brings ...

  22. The Enduring Fascination and Scientific Value of Dinosaurs

    The Scientific Significance of Dinosaurs. Why study dinosaurs? Well, first off, they teach us a ton about science. Dinosaurs were around for about 165 million years, during which a lot of stuff changed on Earth. By looking at their fossils, scientists can learn how life evolved. Take the Archaeopteryx, for example.

  23. Challenges of Science and Technology

    Essay Example: Nowadays, science and tech are super important in shaping our world and pushing things forward. But, even with all the good stuff they bring, there are a bunch of problems we need to tackle. ... Also, job automation due to tech advancements is a big worry. A report by the [Organisation for Economic Co-operation and Development ...

  24. Literature and Science: Matthew Arnold's Analysis

    Essay Example: Introduction Matthew Arnold, a big name in Victorian poetry and culture, had a lot to say about how literature and science relate. In the 1880s, he gave a lecture called "Literature and Science" that dug into the back-and-forth between these two fields. Arnold's thoughts came

  25. GPT-fabricated scientific papers on Google Scholar: Key features

    A sample of scientific papers with signs of GPT-use found on Google Scholar was retrieved, downloaded, and analyzed using a combination of qualitative coding and descriptive statistics. All papers contained at least one of two common phrases returned by conversational agents that use large language models (LLM) like OpenAI's ChatGPT.

  26. Book Review Article

    German research reports; International research reports; Conference reports; ... Book Review Essay | 2000. digital version 109 Book Reviews | 2000. digital version ... Science/Technology 12:30-22:00 Patents and Standards closed ...