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

  • Privacy Policy

Research Method

Home » Research Report – Example, Writing Guide and Types

Research Report – Example, Writing Guide and Types

Table of Contents

Research Report

Research Report

Definition:

Research Report is a written document that presents the results of a research project or study, including the research question, methodology, results, and conclusions, in a clear and objective manner.

The purpose of a research report is to communicate the findings of the research to the intended audience, which could be other researchers, stakeholders, or the general public.

Components of Research Report

Components of Research Report are as follows:

Introduction

The introduction sets the stage for the research report and provides a brief overview of the research question or problem being investigated. It should include a clear statement of the purpose of the study and its significance or relevance to the field of research. It may also provide background information or a literature review to help contextualize the research.

Literature Review

The literature review provides a critical analysis and synthesis of the existing research and scholarship relevant to the research question or problem. It should identify the gaps, inconsistencies, and contradictions in the literature and show how the current study addresses these issues. The literature review also establishes the theoretical framework or conceptual model that guides the research.

Methodology

The methodology section describes the research design, methods, and procedures used to collect and analyze data. It should include information on the sample or participants, data collection instruments, data collection procedures, and data analysis techniques. The methodology should be clear and detailed enough to allow other researchers to replicate the study.

The results section presents the findings of the study in a clear and objective manner. It should provide a detailed description of the data and statistics used to answer the research question or test the hypothesis. Tables, graphs, and figures may be included to help visualize the data and illustrate the key findings.

The discussion section interprets the results of the study and explains their significance or relevance to the research question or problem. It should also compare the current findings with those of previous studies and identify the implications for future research or practice. The discussion should be based on the results presented in the previous section and should avoid speculation or unfounded conclusions.

The conclusion summarizes the key findings of the study and restates the main argument or thesis presented in the introduction. It should also provide a brief overview of the contributions of the study to the field of research and the implications for practice or policy.

The references section lists all the sources cited in the research report, following a specific citation style, such as APA or MLA.

The appendices section includes any additional material, such as data tables, figures, or instruments used in the study, that could not be included in the main text due to space limitations.

Types of Research Report

Types of Research Report are as follows:

Thesis is a type of research report. A thesis is a long-form research document that presents the findings and conclusions of an original research study conducted by a student as part of a graduate or postgraduate program. It is typically written by a student pursuing a higher degree, such as a Master’s or Doctoral degree, although it can also be written by researchers or scholars in other fields.

Research Paper

Research paper is a type of research report. A research paper is a document that presents the results of a research study or investigation. Research papers can be written in a variety of fields, including science, social science, humanities, and business. They typically follow a standard format that includes an introduction, literature review, methodology, results, discussion, and conclusion sections.

Technical Report

A technical report is a detailed report that provides information about a specific technical or scientific problem or project. Technical reports are often used in engineering, science, and other technical fields to document research and development work.

Progress Report

A progress report provides an update on the progress of a research project or program over a specific period of time. Progress reports are typically used to communicate the status of a project to stakeholders, funders, or project managers.

Feasibility Report

A feasibility report assesses the feasibility of a proposed project or plan, providing an analysis of the potential risks, benefits, and costs associated with the project. Feasibility reports are often used in business, engineering, and other fields to determine the viability of a project before it is undertaken.

Field Report

A field report documents observations and findings from fieldwork, which is research conducted in the natural environment or setting. Field reports are often used in anthropology, ecology, and other social and natural sciences.

Experimental Report

An experimental report documents the results of a scientific experiment, including the hypothesis, methods, results, and conclusions. Experimental reports are often used in biology, chemistry, and other sciences to communicate the results of laboratory experiments.

Case Study Report

A case study report provides an in-depth analysis of a specific case or situation, often used in psychology, social work, and other fields to document and understand complex cases or phenomena.

Literature Review Report

A literature review report synthesizes and summarizes existing research on a specific topic, providing an overview of the current state of knowledge on the subject. Literature review reports are often used in social sciences, education, and other fields to identify gaps in the literature and guide future research.

Research Report Example

Following is a Research Report Example sample for Students:

Title: The Impact of Social Media on Academic Performance among High School Students

This study aims to investigate the relationship between social media use and academic performance among high school students. The study utilized a quantitative research design, which involved a survey questionnaire administered to a sample of 200 high school students. The findings indicate that there is a negative correlation between social media use and academic performance, suggesting that excessive social media use can lead to poor academic performance among high school students. The results of this study have important implications for educators, parents, and policymakers, as they highlight the need for strategies that can help students balance their social media use and academic responsibilities.

Introduction:

Social media has become an integral part of the lives of high school students. With the widespread use of social media platforms such as Facebook, Twitter, Instagram, and Snapchat, students can connect with friends, share photos and videos, and engage in discussions on a range of topics. While social media offers many benefits, concerns have been raised about its impact on academic performance. Many studies have found a negative correlation between social media use and academic performance among high school students (Kirschner & Karpinski, 2010; Paul, Baker, & Cochran, 2012).

Given the growing importance of social media in the lives of high school students, it is important to investigate its impact on academic performance. This study aims to address this gap by examining the relationship between social media use and academic performance among high school students.

Methodology:

The study utilized a quantitative research design, which involved a survey questionnaire administered to a sample of 200 high school students. The questionnaire was developed based on previous studies and was designed to measure the frequency and duration of social media use, as well as academic performance.

The participants were selected using a convenience sampling technique, and the survey questionnaire was distributed in the classroom during regular school hours. The data collected were analyzed using descriptive statistics and correlation analysis.

The findings indicate that the majority of high school students use social media platforms on a daily basis, with Facebook being the most popular platform. The results also show a negative correlation between social media use and academic performance, suggesting that excessive social media use can lead to poor academic performance among high school students.

Discussion:

The results of this study have important implications for educators, parents, and policymakers. The negative correlation between social media use and academic performance suggests that strategies should be put in place to help students balance their social media use and academic responsibilities. For example, educators could incorporate social media into their teaching strategies to engage students and enhance learning. Parents could limit their children’s social media use and encourage them to prioritize their academic responsibilities. Policymakers could develop guidelines and policies to regulate social media use among high school students.

Conclusion:

In conclusion, this study provides evidence of the negative impact of social media on academic performance among high school students. The findings highlight the need for strategies that can help students balance their social media use and academic responsibilities. Further research is needed to explore the specific mechanisms by which social media use affects academic performance and to develop effective strategies for addressing this issue.

Limitations:

One limitation of this study is the use of convenience sampling, which limits the generalizability of the findings to other populations. Future studies should use random sampling techniques to increase the representativeness of the sample. Another limitation is the use of self-reported measures, which may be subject to social desirability bias. Future studies could use objective measures of social media use and academic performance, such as tracking software and school records.

Implications:

The findings of this study have important implications for educators, parents, and policymakers. Educators could incorporate social media into their teaching strategies to engage students and enhance learning. For example, teachers could use social media platforms to share relevant educational resources and facilitate online discussions. Parents could limit their children’s social media use and encourage them to prioritize their academic responsibilities. They could also engage in open communication with their children to understand their social media use and its impact on their academic performance. Policymakers could develop guidelines and policies to regulate social media use among high school students. For example, schools could implement social media policies that restrict access during class time and encourage responsible use.

References:

  • Kirschner, P. A., & Karpinski, A. C. (2010). Facebook® and academic performance. Computers in Human Behavior, 26(6), 1237-1245.
  • Paul, J. A., Baker, H. M., & Cochran, J. D. (2012). Effect of online social networking on student academic performance. Journal of the Research Center for Educational Technology, 8(1), 1-19.
  • Pantic, I. (2014). Online social networking and mental health. Cyberpsychology, Behavior, and Social Networking, 17(10), 652-657.
  • Rosen, L. D., Carrier, L. M., & Cheever, N. A. (2013). Facebook and texting made me do it: Media-induced task-switching while studying. Computers in Human Behavior, 29(3), 948-958.

Note*: Above mention, Example is just a sample for the students’ guide. Do not directly copy and paste as your College or University assignment. Kindly do some research and Write your own.

Applications of Research Report

Research reports have many applications, including:

  • Communicating research findings: The primary application of a research report is to communicate the results of a study to other researchers, stakeholders, or the general public. The report serves as a way to share new knowledge, insights, and discoveries with others in the field.
  • Informing policy and practice : Research reports can inform policy and practice by providing evidence-based recommendations for decision-makers. For example, a research report on the effectiveness of a new drug could inform regulatory agencies in their decision-making process.
  • Supporting further research: Research reports can provide a foundation for further research in a particular area. Other researchers may use the findings and methodology of a report to develop new research questions or to build on existing research.
  • Evaluating programs and interventions : Research reports can be used to evaluate the effectiveness of programs and interventions in achieving their intended outcomes. For example, a research report on a new educational program could provide evidence of its impact on student performance.
  • Demonstrating impact : Research reports can be used to demonstrate the impact of research funding or to evaluate the success of research projects. By presenting the findings and outcomes of a study, research reports can show the value of research to funders and stakeholders.
  • Enhancing professional development : Research reports can be used to enhance professional development by providing a source of information and learning for researchers and practitioners in a particular field. For example, a research report on a new teaching methodology could provide insights and ideas for educators to incorporate into their own practice.

How to write Research Report

Here are some steps you can follow to write a research report:

  • Identify the research question: The first step in writing a research report is to identify your research question. This will help you focus your research and organize your findings.
  • Conduct research : Once you have identified your research question, you will need to conduct research to gather relevant data and information. This can involve conducting experiments, reviewing literature, or analyzing data.
  • Organize your findings: Once you have gathered all of your data, you will need to organize your findings in a way that is clear and understandable. This can involve creating tables, graphs, or charts to illustrate your results.
  • Write the report: Once you have organized your findings, you can begin writing the report. Start with an introduction that provides background information and explains the purpose of your research. Next, provide a detailed description of your research methods and findings. Finally, summarize your results and draw conclusions based on your findings.
  • Proofread and edit: After you have written your report, be sure to proofread and edit it carefully. Check for grammar and spelling errors, and make sure that your report is well-organized and easy to read.
  • Include a reference list: Be sure to include a list of references that you used in your research. This will give credit to your sources and allow readers to further explore the topic if they choose.
  • Format your report: Finally, format your report according to the guidelines provided by your instructor or organization. This may include formatting requirements for headings, margins, fonts, and spacing.

Purpose of Research Report

The purpose of a research report is to communicate the results of a research study to a specific audience, such as peers in the same field, stakeholders, or the general public. The report provides a detailed description of the research methods, findings, and conclusions.

Some common purposes of a research report include:

  • Sharing knowledge: A research report allows researchers to share their findings and knowledge with others in their field. This helps to advance the field and improve the understanding of a particular topic.
  • Identifying trends: A research report can identify trends and patterns in data, which can help guide future research and inform decision-making.
  • Addressing problems: A research report can provide insights into problems or issues and suggest solutions or recommendations for addressing them.
  • Evaluating programs or interventions : A research report can evaluate the effectiveness of programs or interventions, which can inform decision-making about whether to continue, modify, or discontinue them.
  • Meeting regulatory requirements: In some fields, research reports are required to meet regulatory requirements, such as in the case of drug trials or environmental impact studies.

When to Write Research Report

A research report should be written after completing the research study. This includes collecting data, analyzing the results, and drawing conclusions based on the findings. Once the research is complete, the report should be written in a timely manner while the information is still fresh in the researcher’s mind.

In academic settings, research reports are often required as part of coursework or as part of a thesis or dissertation. In this case, the report should be written according to the guidelines provided by the instructor or institution.

In other settings, such as in industry or government, research reports may be required to inform decision-making or to comply with regulatory requirements. In these cases, the report should be written as soon as possible after the research is completed in order to inform decision-making in a timely manner.

Overall, the timing of when to write a research report depends on the purpose of the research, the expectations of the audience, and any regulatory requirements that need to be met. However, it is important to complete the report in a timely manner while the information is still fresh in the researcher’s mind.

Characteristics of Research Report

There are several characteristics of a research report that distinguish it from other types of writing. These characteristics include:

  • Objective: A research report should be written in an objective and unbiased manner. It should present the facts and findings of the research study without any personal opinions or biases.
  • Systematic: A research report should be written in a systematic manner. It should follow a clear and logical structure, and the information should be presented in a way that is easy to understand and follow.
  • Detailed: A research report should be detailed and comprehensive. It should provide a thorough description of the research methods, results, and conclusions.
  • Accurate : A research report should be accurate and based on sound research methods. The findings and conclusions should be supported by data and evidence.
  • Organized: A research report should be well-organized. It should include headings and subheadings to help the reader navigate the report and understand the main points.
  • Clear and concise: A research report should be written in clear and concise language. The information should be presented in a way that is easy to understand, and unnecessary jargon should be avoided.
  • Citations and references: A research report should include citations and references to support the findings and conclusions. This helps to give credit to other researchers and to provide readers with the opportunity to further explore the topic.

Advantages of Research Report

Research reports have several advantages, including:

  • Communicating research findings: Research reports allow researchers to communicate their findings to a wider audience, including other researchers, stakeholders, and the general public. This helps to disseminate knowledge and advance the understanding of a particular topic.
  • Providing evidence for decision-making : Research reports can provide evidence to inform decision-making, such as in the case of policy-making, program planning, or product development. The findings and conclusions can help guide decisions and improve outcomes.
  • Supporting further research: Research reports can provide a foundation for further research on a particular topic. Other researchers can build on the findings and conclusions of the report, which can lead to further discoveries and advancements in the field.
  • Demonstrating expertise: Research reports can demonstrate the expertise of the researchers and their ability to conduct rigorous and high-quality research. This can be important for securing funding, promotions, and other professional opportunities.
  • Meeting regulatory requirements: In some fields, research reports are required to meet regulatory requirements, such as in the case of drug trials or environmental impact studies. Producing a high-quality research report can help ensure compliance with these requirements.

Limitations of Research Report

Despite their advantages, research reports also have some limitations, including:

  • Time-consuming: Conducting research and writing a report can be a time-consuming process, particularly for large-scale studies. This can limit the frequency and speed of producing research reports.
  • Expensive: Conducting research and producing a report can be expensive, particularly for studies that require specialized equipment, personnel, or data. This can limit the scope and feasibility of some research studies.
  • Limited generalizability: Research studies often focus on a specific population or context, which can limit the generalizability of the findings to other populations or contexts.
  • Potential bias : Researchers may have biases or conflicts of interest that can influence the findings and conclusions of the research study. Additionally, participants may also have biases or may not be representative of the larger population, which can limit the validity and reliability of the findings.
  • Accessibility: Research reports may be written in technical or academic language, which can limit their accessibility to a wider audience. Additionally, some research may be behind paywalls or require specialized access, which can limit the ability of others to read and use the findings.

About the author

' src=

Muhammad Hassan

Researcher, Academic Writer, Web developer

You may also like

References in Research

References in Research – Types, Examples and...

Research Objectives

Research Objectives – Types, Examples and...

Research Results

Research Results Section – Writing Guide and...

Evaluating Research

Evaluating Research – Process, Examples and...

Future Research

Future Research – Thesis Guide

Research Contribution

Research Contribution – Thesis Guide

Have a language expert improve your writing

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

  • Knowledge Base
  • Research paper

How to Write a Research Paper | A Beginner's Guide

A research paper is a piece of academic writing that provides analysis, interpretation, and argument based on in-depth independent research.

Research papers are similar to academic essays , but they are usually longer and more detailed assignments, designed to assess not only your writing skills but also your skills in scholarly research. Writing a research paper requires you to demonstrate a strong knowledge of your topic, engage with a variety of sources, and make an original contribution to the debate.

This step-by-step guide takes you through the entire writing process, from understanding your assignment to proofreading your final draft.

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

Understand the assignment, choose a research paper topic, conduct preliminary research, develop a thesis statement, create a research paper outline, write a first draft of the research paper, write the introduction, write a compelling body of text, write the conclusion, the second draft, the revision process, research paper checklist, free lecture slides.

Completing a research paper successfully means accomplishing the specific tasks set out for you. Before you start, make sure you thoroughly understanding the assignment task sheet:

  • Read it carefully, looking for anything confusing you might need to clarify with your professor.
  • Identify the assignment goal, deadline, length specifications, formatting, and submission method.
  • Make a bulleted list of the key points, then go back and cross completed items off as you’re writing.

Carefully consider your timeframe and word limit: be realistic, and plan enough time to research, write, and edit.

Here's why students love Scribbr's proofreading services

Discover proofreading & editing

There are many ways to generate an idea for a research paper, from brainstorming with pen and paper to talking it through with a fellow student or professor.

You can try free writing, which involves taking a broad topic and writing continuously for two or three minutes to identify absolutely anything relevant that could be interesting.

You can also gain inspiration from other research. The discussion or recommendations sections of research papers often include ideas for other specific topics that require further examination.

Once you have a broad subject area, narrow it down to choose a topic that interests you, m eets the criteria of your assignment, and i s possible to research. Aim for ideas that are both original and specific:

  • A paper following the chronology of World War II would not be original or specific enough.
  • A paper on the experience of Danish citizens living close to the German border during World War II would be specific and could be original enough.

Note any discussions that seem important to the topic, and try to find an issue that you can focus your paper around. Use a variety of sources , including journals, books, and reliable websites, to ensure you do not miss anything glaring.

Do not only verify the ideas you have in mind, but look for sources that contradict your point of view.

  • Is there anything people seem to overlook in the sources you research?
  • Are there any heated debates you can address?
  • Do you have a unique take on your topic?
  • Have there been some recent developments that build on the extant research?

In this stage, you might find it helpful to formulate some research questions to help guide you. To write research questions, try to finish the following sentence: “I want to know how/what/why…”

A thesis statement is a statement of your central argument — it establishes the purpose and position of your paper. If you started with a research question, the thesis statement should answer it. It should also show what evidence and reasoning you’ll use to support that answer.

The thesis statement should be concise, contentious, and coherent. That means it should briefly summarize your argument in a sentence or two, make a claim that requires further evidence or analysis, and make a coherent point that relates to every part of the paper.

You will probably revise and refine the thesis statement as you do more research, but it can serve as a guide throughout the writing process. Every paragraph should aim to support and develop this central claim.

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

how to write scientific research report

A research paper outline is essentially a list of the key topics, arguments, and evidence you want to include, divided into sections with headings so that you know roughly what the paper will look like before you start writing.

A structure outline can help make the writing process much more efficient, so it’s worth dedicating some time to create one.

Your first draft won’t be perfect — you can polish later on. Your priorities at this stage are as follows:

  • Maintaining forward momentum — write now, perfect later.
  • Paying attention to clear organization and logical ordering of paragraphs and sentences, which will help when you come to the second draft.
  • Expressing your ideas as clearly as possible, so you know what you were trying to say when you come back to the text.

You do not need to start by writing the introduction. Begin where it feels most natural for you — some prefer to finish the most difficult sections first, while others choose to start with the easiest part. If you created an outline, use it as a map while you work.

Do not delete large sections of text. If you begin to dislike something you have written or find it doesn’t quite fit, move it to a different document, but don’t lose it completely — you never know if it might come in useful later.

Paragraph structure

Paragraphs are the basic building blocks of research papers. Each one should focus on a single claim or idea that helps to establish the overall argument or purpose of the paper.

Example paragraph

George Orwell’s 1946 essay “Politics and the English Language” has had an enduring impact on thought about the relationship between politics and language. This impact is particularly obvious in light of the various critical review articles that have recently referenced the essay. For example, consider Mark Falcoff’s 2009 article in The National Review Online, “The Perversion of Language; or, Orwell Revisited,” in which he analyzes several common words (“activist,” “civil-rights leader,” “diversity,” and more). Falcoff’s close analysis of the ambiguity built into political language intentionally mirrors Orwell’s own point-by-point analysis of the political language of his day. Even 63 years after its publication, Orwell’s essay is emulated by contemporary thinkers.

Citing sources

It’s also important to keep track of citations at this stage to avoid accidental plagiarism . Each time you use a source, make sure to take note of where the information came from.

You can use our free citation generators to automatically create citations and save your reference list as you go.

APA Citation Generator MLA Citation Generator

The research paper introduction should address three questions: What, why, and how? After finishing the introduction, the reader should know what the paper is about, why it is worth reading, and how you’ll build your arguments.

What? Be specific about the topic of the paper, introduce the background, and define key terms or concepts.

Why? This is the most important, but also the most difficult, part of the introduction. Try to provide brief answers to the following questions: What new material or insight are you offering? What important issues does your essay help define or answer?

How? To let the reader know what to expect from the rest of the paper, the introduction should include a “map” of what will be discussed, briefly presenting the key elements of the paper in chronological order.

The major struggle faced by most writers is how to organize the information presented in the paper, which is one reason an outline is so useful. However, remember that the outline is only a guide and, when writing, you can be flexible with the order in which the information and arguments are presented.

One way to stay on track is to use your thesis statement and topic sentences . Check:

  • topic sentences against the thesis statement;
  • topic sentences against each other, for similarities and logical ordering;
  • and each sentence against the topic sentence of that paragraph.

Be aware of paragraphs that seem to cover the same things. If two paragraphs discuss something similar, they must approach that topic in different ways. Aim to create smooth transitions between sentences, paragraphs, and sections.

The research paper conclusion is designed to help your reader out of the paper’s argument, giving them a sense of finality.

Trace the course of the paper, emphasizing how it all comes together to prove your thesis statement. Give the paper a sense of finality by making sure the reader understands how you’ve settled the issues raised in the introduction.

You might also discuss the more general consequences of the argument, outline what the paper offers to future students of the topic, and suggest any questions the paper’s argument raises but cannot or does not try to answer.

You should not :

  • Offer new arguments or essential information
  • Take up any more space than necessary
  • Begin with stock phrases that signal you are ending the paper (e.g. “In conclusion”)

There are four main considerations when it comes to the second draft.

  • Check how your vision of the paper lines up with the first draft and, more importantly, that your paper still answers the assignment.
  • Identify any assumptions that might require (more substantial) justification, keeping your reader’s perspective foremost in mind. Remove these points if you cannot substantiate them further.
  • Be open to rearranging your ideas. Check whether any sections feel out of place and whether your ideas could be better organized.
  • If you find that old ideas do not fit as well as you anticipated, you should cut them out or condense them. You might also find that new and well-suited ideas occurred to you during the writing of the first draft — now is the time to make them part of the paper.

The goal during the revision and proofreading process is to ensure you have completed all the necessary tasks and that the paper is as well-articulated as possible. You can speed up the proofreading process by using the AI proofreader .

Global concerns

  • Confirm that your paper completes every task specified in your assignment sheet.
  • Check for logical organization and flow of paragraphs.
  • Check paragraphs against the introduction and thesis statement.

Fine-grained details

Check the content of each paragraph, making sure that:

  • each sentence helps support the topic sentence.
  • no unnecessary or irrelevant information is present.
  • all technical terms your audience might not know are identified.

Next, think about sentence structure , grammatical errors, and formatting . Check that you have correctly used transition words and phrases to show the connections between your ideas. Look for typos, cut unnecessary words, and check for consistency in aspects such as heading formatting and spellings .

Finally, you need to make sure your paper is correctly formatted according to the rules of the citation style you are using. For example, you might need to include an MLA heading  or create an APA title page .

Scribbr’s professional editors can help with the revision process with our award-winning proofreading services.

Discover our paper editing service

Checklist: Research paper

I have followed all instructions in the assignment sheet.

My introduction presents my topic in an engaging way and provides necessary background information.

My introduction presents a clear, focused research problem and/or thesis statement .

My paper is logically organized using paragraphs and (if relevant) section headings .

Each paragraph is clearly focused on one central idea, expressed in a clear topic sentence .

Each paragraph is relevant to my research problem or thesis statement.

I have used appropriate transitions  to clarify the connections between sections, paragraphs, and sentences.

My conclusion provides a concise answer to the research question or emphasizes how the thesis has been supported.

My conclusion shows how my research has contributed to knowledge or understanding of my topic.

My conclusion does not present any new points or information essential to my argument.

I have provided an in-text citation every time I refer to ideas or information from a source.

I have included a reference list at the end of my paper, consistently formatted according to a specific citation style .

I have thoroughly revised my paper and addressed any feedback from my professor or supervisor.

I have followed all formatting guidelines (page numbers, headers, spacing, etc.).

You've written a great paper. Make sure it's perfect with the help of a Scribbr editor!

Open Google Slides Download PowerPoint

Is this article helpful?

Other students also liked.

  • Writing a Research Paper Introduction | Step-by-Step Guide
  • Writing a Research Paper Conclusion | Step-by-Step Guide
  • Research Paper Format | APA, MLA, & Chicago Templates

More interesting articles

  • Academic Paragraph Structure | Step-by-Step Guide & Examples
  • Checklist: Writing a Great Research Paper
  • How to Create a Structured Research Paper Outline | Example
  • How to Write a Discussion Section | Tips & Examples
  • How to Write Recommendations in Research | Examples & Tips
  • How to Write Topic Sentences | 4 Steps, Examples & Purpose
  • Research Paper Appendix | Example & Templates
  • Research Paper Damage Control | Managing a Broken Argument
  • What Is a Theoretical Framework? | Guide to Organizing

"I thought AI Proofreading was useless but.."

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

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
  • INTRODUCTION

Writing a "good" results section

Figures and Captions in Lab Reports

"Results Checklist" from: How to Write a Good Scientific Paper. Chris A. Mack. SPIE. 2018.

Additional tips for results sections.

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

This is the core of the paper. Don't start the results sections with methods you left out of the Materials and Methods section. You need to give an overall description of the experiments and present the data you found.

  • Factual statements supported by evidence. Short and sweet without excess words
  • Present representative data rather than endlessly repetitive data
  • Discuss variables only if they had an effect (positive or negative)
  • Use meaningful statistics
  • Avoid redundancy. If it is in the tables or captions you may not need to repeat it

A short article by Dr. Brett Couch and Dr. Deena Wassenberg, Biology Program, University of Minnesota

  • 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 discussed;  presenting results in chronological order rather than logical order; ignoring results that do not support the conclusions; 
  • Number tables and figures separately beginning with 1 (i.e. Table 1, Table 2, Figure 1, etc.).
  • Do not attempt to evaluate the results in this section. Report only what you found; hold all discussion of the significance of the results for the Discussion section.
  • It is not necessary to describe every step of your statistical analyses. Scientists understand all about null hypotheses, rejection rules, and so forth and do not need to be reminded of them. Just say something like, "Honeybees did not use the flowers in proportion to their availability (X2 = 7.9, p<0.05, d.f.= 4, chi-square test)." Likewise, cite tables and figures without describing in detail how the data were manipulated. Explanations of this sort should appear in a legend or caption written on the same page as the figure or table.
  • You must refer in the text to each figure or table you include in your paper.
  • Tables generally should report summary-level data, such as means ± standard deviations, rather than all your raw data.  A long list of all your individual observations will mean much less than a few concise, easy-to-read tables or figures that bring out the main findings of your study.  
  • Only use a figure (graph) when the data lend themselves to a good visual representation.  Avoid using figures that show too many variables or trends at once, because they can be hard to understand.

From:  https://writingcenter.gmu.edu/guides/imrad-results-discussion

  • << Previous: METHODS
  • Next: DISCUSSION >>
  • 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.

How to Write a Scientific Paper

  • First Online: 01 October 2023

Cite this chapter

how to write scientific research report

  • Michael J. Curtis 4  

1133 Accesses

A scientific paper is a report of research, prepared from the investigator’s experimental findings, and intended to contribute to knowledge. The publication process is predicated by the content (the data) and the requirements of the publication (which is normally a journal that exists in paper form and/or online) whom you wish to publish your work. The requirement of the investigator is to map their data to the structure required by the journal. It is essential, therefore, to read and understand the journal’s Instructions to Authors. This means the subject matter must map to the journal’s scope, and the manuscript structure must map to the journal’s needs. If the investigator can do this, the manuscript may then be sent by the journal for peer review. This means there is no single formula for creating a publishable item; publishability depends on the content, presentation and the requirements of the publication . As a journal editor, I frequently reject items without peer review because they are not in scope (‘off topic’) or do not follow key construction requirements (e.g., Results and Discussion presented as a single section rather than as separate sections). The author must decide whether their research is appropriate for their chosen journal. In many cases there may not be sufficient data, or data of an appropriate type to justify a publication in all but the least discerning of journals. The investigator may learn lessons from this. Here I provide guidance on navigating the process of writing a scientific paper.

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

Access this chapter

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
  • Available as EPUB and PDF
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

how to write scientific research report

Writing and publishing a scientific paper

how to write scientific research report

How to Write and Publish a Research Paper for a Peer-Reviewed Journal

how to write scientific research report

The Point Is…to Publish?

Curtis MJ, Hancox JC, Farkas A, Wainwright CL, Stables CL, Saint DA, Clements-Jewery H, Lambiase PD, Billman GE, Janse MJ, Pugsley MK, Ng GN, Roden DM, Camm AJ, Walker MJA (2013) The Lambeth Conventions (II): guidelines for the study of animal and human ventricular and supraventricular arrhythmias. Pharmacol Ther 139:213–248

Article   CAS   PubMed   Google Scholar  

Curtis MJ, Bond RA, Spina D, Ahluwalia A, Alexander SPA, Giembycz MA, Gilchrist A, Hoyer D, Insel P, Izzo AA, Lawrence AJ, MacEwan DJ, Moon LDF, Wonnacott S, Weston AH, McGrath JC (2015) Experimental design and analysis and their reporting: new guidance for publication in BJP. Br J Pharmacol 172:2671–2674

Article   PubMed Central   Google Scholar  

Curtis MJ, Alexander S, Cirino G, Docherty JR, George CH, Giembycz MA, Hoyer D, Insel PA, Izzo AA, Ji Y, MacEwan DA, Sobey CG, Stanford SC, Teixeira MM, Wonnacott S, Ahluwalia A (2018) Experimental design and analysis and their reporting II: updated and simplified guidance for authors and peer reviewers. Br J Pharmacol 175:987–993

Article   CAS   PubMed   PubMed Central   Google Scholar  

Wilder CDE, Pavlaki N, Dursun T, Gyimah P, Caldwell-Dunn E, Ranieri A, Lewis HR, Curtis MJ (2018) Facilitation of ischaemia-induced ventricular fibrillation by catecholamines is mediated by β1 and β2 agonism in the rat heart in vitro. Br J Pharmacol 175:1669–1690

Curtis MJ, Walker MJA (1988) Quantification of arrhythmias using scoring systems: an examination of seven scores in an in vivo model of regional myocardial ischaemia. Cardiovasc Res 22:656–665

Williams M, Mullane K, Curtis MJ (2018) Addressing reproducibility: peer review, impact factors, checklists, guidelines, and reproducibility initiatives. In: Williams M, Mullane K, Curtis MJ (eds) Research in the biomedical sciences. Elsevier, New York, pp 197–306

Chapter   Google Scholar  

Rees SA, Curtis MJ (1995) A pharmacological analysis in rat of the role of the ATP-sensitive potassium channel as a target for antifibrillatory intervention in acute myocardial ischaemia. J Cardiovasc Pharmacol 26:319–327

Article   Google Scholar  

Download references

Acknowledgements

Conflict of interest statement, author information, authors and affiliations.

Cardiovascular Division, Faculty of Life Sciences and Medicine, School of Cardiovascular Medicine & Sciences, King’s College London, Rayne Institute, St Thomas’ Hospital, London, UK

Michael J. Curtis

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Michael J. Curtis .

Editor information

Editors and affiliations.

Retired Senior Expert Pharmacologist at the Office of Cardiology, Hematology, Endocrinology, and Nephrology, Center for Drug Evaluation and Research, US Food and Drug Administration, Silver Spring, MD, USA

Gowraganahalli Jagadeesh

Professor & Director, Research Training and Publications, The Office of Research and Development, Periyar Maniammai Institute of Science & Technology (Deemed to be University), Vallam, Tamil Nadu, India

Pitchai Balakumar

Division Cardiology & Nephrology, Office of Cardiology, Hematology, Endocrinology and Nephrology, Center for Drug Evaluation and Research, US Food and Drug Administration, Silver Spring, MD, USA

Fortunato Senatore

Rights and permissions

Reprints and permissions

Copyright information

© 2023 The Author(s), under exclusive license to Springer Nature Singapore Pte Ltd.

About this chapter

Curtis, M.J. (2023). How to Write a Scientific Paper. In: Jagadeesh, G., Balakumar, P., Senatore, F. (eds) The Quintessence of Basic and Clinical Research and Scientific Publishing. Springer, Singapore. https://doi.org/10.1007/978-981-99-1284-1_41

Download citation

DOI : https://doi.org/10.1007/978-981-99-1284-1_41

Published : 01 October 2023

Publisher Name : Springer, Singapore

Print ISBN : 978-981-99-1283-4

Online ISBN : 978-981-99-1284-1

eBook Packages : Biomedical and Life Sciences Biomedical and Life Sciences (R0)

Share this chapter

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Publish with us

Policies and ethics

  • Find a journal
  • Track your research
  • Your Science & Health Librarians
  • How To Find Articles with Databases
  • Video Learning
  • Artificial Intelligence Tools
  • Industry Reports
  • How To Evaluate Articles
  • Search Tips, General
  • Develop a Research Question
  • How To Read A Scientific Paper
  • How To Interpret Data
  • How To Write A Scientific Paper
  • Teaching Materials
  • Systematic & Evideced-Based Reviews
  • Get More Help

Writing a Scientific Paper or Lab Report

Writing a scientific paper is very similar to writing a lab report. The structure of each is primarily the same, but the purpose of each is different

Lab reports are meant to reflect understanding of the material and learn something new Scientific papers are meant to contribute knowledge to a field of study.

Both are generally broken down into eight sections: title, abstract, introduction, methods, results, discussion, conclusion, and references. 

  • Ex: Determining the Free Chlorine Content of Pool Water"
  • Abstracts are a summary of the research as a whole and should familiarize the reader with the purpose of the research. 
  • Abstracts will always be written last, even though they are the first paragraph of a scientific paper. 
  • Unlike a lab report, all scientific papers will have an abstract.
  • Why was the research done?
  • What problem is being addressed?
  • What results were found?
  • What are the meaning of the results?
  • How is the problem better understood now than before, if at all?

Introduction

  • The introduction of a scientific paper discusses the problem being studied and other theory that is relevant to understanding the findings. 
  • The hypothesis of the experiment and the motivation for the research are stated in this section. 
  • Write the introduction in your own words. Try not to copy from a lab manual or other guidelines. Instead, show comprehension of the research by briefly explaining the problem.
  • Methods and Materials
  • Ex: pipette, graduated cylinder, 1.13mg of Na, 0.67mg Ag
  • List the steps taken as they actually happened during the experiment, not as they were supposed to happen. 
  • If written correctly, another researcher should be able to duplicate the experiment and get the same or very similar results. 
  • In a scientific paper, most often the steps taken during the research are discussed more in length and with more detail than they are in lab reports. 
  • The results show the data that was collected or found during the research. 
  • Explain in words the data that was collected.
  • Tables should be labeled numerically, as "Table 1", "Table 2", etc. Other figures should be labeled numerically as "Figure 1", "Figure 2", etc. 
  • Calculations to understand the data can also be presented in the results. 
  • The discussion section is one of the most important parts of a scientific paper. It analyzes the results of the research and is a discussion of the data. 
  • If any results are unexpected, explain why they are unexpected and how they did or did not effect the data obtained. 
  • Analyze the strengths and weaknesses of the design of the research and compare your results to similar research.
  • If there are any experimental errors, analyze them.
  • Explain your results and discuss them using relevant terms and theories.
  • What do the results indicate?
  • What is the significance of the results?
  • Are there any gaps in knowledge?
  • Are there any new questions that have been raised?
  • The conclusion is a summation of the experiment. It should clearly and concisely state what was learned and its importance.
  • If there is future work that needs to be done, it can be explained in the conclusion.
  • When any outside sources to support a claim or explain background information, those sources must be cited in the references section of the lab report. 
  • Scientific papers will always use outside references. 

Other Useful Sources

  • How to Write a Scientific Article
  • Writing a Scientific Research Article
  • How to Write a Good Scientific Paper
  • << Previous: How To Interpret Data
  • Next: Teaching Materials >>
  • Last Updated: Jul 26, 2024 11:56 AM
  • URL: https://guides.libraries.indiana.edu/STEM

Social media

  • Instagram for Herman B Wells Library
  • Facebook for IU Libraries

Additional resources

Featured databases.

  • Resource available to authorized IU Bloomington users (on or off campus) OneSearch@IU
  • Resource available to authorized IU Bloomington users (on or off campus) Academic Search (EBSCO)
  • Resource available to authorized IU Bloomington users (on or off campus) ERIC (EBSCO)
  • Resource available to authorized IU Bloomington users (on or off campus) Nexis Uni
  • Resource available without restriction HathiTrust Digital Library
  • Databases A-Z
  • Resource available to authorized IU Bloomington users (on or off campus) Google Scholar
  • Resource available to authorized IU Bloomington users (on or off campus) JSTOR
  • Resource available to authorized IU Bloomington users (on or off campus) Web of Science
  • Resource available to authorized IU Bloomington users (on or off campus) Scopus
  • Resource available to authorized IU Bloomington users (on or off campus) WorldCat

IU Libraries

  • Diversity Resources
  • About IU Libraries
  • Alumni & Friends
  • Departments & Staff
  • Jobs & Libraries HR
  • Intranet (Staff)
  • IUL site admin

U.S. flag

An official website of the United States government

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

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

  • Publications
  • Account settings

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

  • Advanced Search
  • Journal List

Logo of zeb

A Guide to Writing a Scientific Paper: A Focus on High School Through Graduate Level Student Research

Renee a. hesselbach.

1 NIEHS Children's Environmental Health Sciences Core Center, University of Wisconsin—Milwaukee, Milwaukee, Wisconsin.

David H. Petering

2 Department of Chemistry and Biochemistry, University of Wisconsin—Milwaukee, Milwaukee, Wisconsin.

Craig A. Berg

3 Curriculum and Instruction, University of Wisconsin—Milwaukee, Milwaukee, Wisconsin.

Henry Tomasiewicz

Daniel weber.

This article presents a detailed guide for high school through graduate level instructors that leads students to write effective and well-organized scientific papers. Interesting research emerges from the ability to ask questions, define problems, design experiments, analyze and interpret data, and make critical connections. This process is incomplete, unless new results are communicated to others because science fundamentally requires peer review and criticism to validate or discard proposed new knowledge. Thus, a concise and clearly written research paper is a critical step in the scientific process and is important for young researchers as they are mastering how to express scientific concepts and understanding. Moreover, learning to write a research paper provides a tool to improve science literacy as indicated in the National Research Council's National Science Education Standards (1996), and A Framework for K–12 Science Education (2011), the underlying foundation for the Next Generation Science Standards currently being developed. Background information explains the importance of peer review and communicating results, along with details of each critical component, the Abstract, Introduction, Methods, Results , and Discussion . Specific steps essential to helping students write clear and coherent research papers that follow a logical format, use effective communication, and develop scientific inquiry are described.

Introduction

A key part of the scientific process is communication of original results to others so that one's discoveries are passed along to the scientific community and the public for awareness and scrutiny. 1 – 3 Communication to other scientists ensures that new findings become part of a growing body of publicly available knowledge that informs how we understand the world around us. 2 It is also what fuels further research as other scientists incorporate novel findings into their thinking and experiments.

Depending upon the researcher's position, intent, and needs, communication can take different forms. The gold standard is writing scientific papers that describe original research in such a way that other scientists will be able to repeat it or to use it as a basis for their studies. 1 For some, it is expected that such articles will be published in scientific journals after they have been peer reviewed and accepted for publication. Scientists must submit their articles for examination by other scientists familiar with the area of research, who decide whether the work was conducted properly and whether the results add to the knowledge base and are conveyed well enough to merit publication. 2 If a manuscript passes the scrutiny of peer-review, it has the potential to be published. 1 For others, such as for high school or undergraduate students, publishing a research paper may not be the ultimate goal. However, regardless of whether an article is to be submitted for publication, peer review is an important step in this process. For student researchers, writing a well-organized research paper is a key step in learning how to express understanding, make critical connections, summarize data, and effectively communicate results, which are important goals for improving science literacy of the National Research Council's National Science Education Standards, 4 and A Framework for K–12 Science Education, 5 and the Next Generation Science Standards 6 currently being developed and described in The NSTA Reader's Guide to A Framework for K–12 Science Education. 7 Table 1 depicts the key skills students should develop as part of the Science as Inquiry Content Standard. Table 2 illustrates the central goals of A Framework for K–12 Science Education Scientific and Engineering Practices Dimension.

Key Skills of the Science as Inquiry National Science Education Content Standard

Identify questions and concepts that guide scientific investigation
Design and conduct scientific investigations
Use technology and mathematics to improve investigations and communications
Formulate and revise scientific explanations and models using logic and evidence
Recognize and analyze alternative explanations and models
Communicate and defend a scientific argument

National Research Council (1996).

Important Practices of A Framework for K–12 Science Education Scientific and Engineering Practices Dimension

Asking questions and defining problems
Developing and using models
Planning and carrying out investigations
Analyzing and interpreting data
Using mathematics and computational thinking
Constructing explanations and designing solutions
Engaging in argument from evidence
Obtaining, evaluating, and communicating information

National Research Council (2011).

Scientific papers based on experimentation typically include five predominant sections: Abstract, Introduction, Methods, Results, and Discussion . This structure is a widely accepted approach to writing a research paper, and has specific sections that parallel the scientific method. Following this structure allows the scientist to tell a clear, coherent story in a logical format, essential to effective communication. 1 , 2 In addition, using a standardized format allows the reader to find specific information quickly and easily. While readers may not have time to read the entire research paper, the predictable format allows them to focus on specific sections such as the Abstract , Introduction , and Discussion sections. Therefore, it is critical that information be placed in the appropriate and logical section of the report. 3

Guidelines for Writing a Primary Research Article

The Title sends an important message to the reader about the purpose of the paper. For example, Ethanol Effects on the Developing Zebrafish: Neurobehavior and Skeletal Morphogenesis 8 tells the reader key information about the content of the research paper. Also, an appropriate and descriptive title captures the attention of the reader. When composing the Title , students should include either the aim or conclusion of the research, the subject, and possibly the independent or dependent variables. Often, the title is created after the body of the article has been written, so that it accurately reflects the purpose and content of the article. 1 , 3

The Abstract provides a short, concise summary of the research described in the body of the article and should be able to stand alone. It provides readers with a quick overview that helps them decide whether the article may be interesting to read. Included in the Abstract are the purpose or primary objectives of the experiment and why they are important, a brief description of the methods and approach used, key findings and the significance of the results, and how this work is different from the work of others. It is important to note that the Abstract briefly explains the implications of the findings, but does not evaluate the conclusions. 1 , 3 Just as with the Title , this section needs to be written carefully and succinctly. Often this section is written last to ensure it accurately reflects the content of the paper. Generally, the optimal length of the Abstract is one paragraph between 200 and 300 words, and does not contain references or abbreviations.

All new research can be categorized by field (e.g., biology, chemistry, physics, geology) and by area within the field (e.g., biology: evolution, ecology, cell biology, anatomy, environmental health). Many areas already contain a large volume of published research. The role of the Introduction is to place the new research within the context of previous studies in the particular field and area, thereby introducing the audience to the research and motivating the audience to continue reading. 1

Usually, the writer begins by describing what is known in the area that directly relates to the subject of the article's research. Clearly, this must be done judiciously; usually there is not room to describe every bit of information that is known. Each statement needs one or more references from the scientific literature that supports its validity. Students must be reminded to cite all references to eliminate the risk of plagiarism. 2 Out of this context, the author then explains what is not known and, therefore, what the article's research seeks to find out. In doing so, the scientist provides the rationale for the research and further develops why this research is important. The final statement in the Introduction should be a clearly worded hypothesis or thesis statement, as well as a brief summary of the findings as they relate to the stated hypothesis. Keep in mind that the details of the experimental findings are presented in the Results section and are aimed at filling the void in our knowledge base that has been pointed out in the Introduction .

Materials and Methods

Research utilizes various accepted methods to obtain the results that are to be shared with others in the scientific community. The quality of the results, therefore, depends completely upon the quality of the methods that are employed and the care with which they are applied. The reader will refer to the Methods section: (a) to become confident that the experiments have been properly done, (b) as the guide for repeating the experiments, and (c) to learn how to do new methods.

It is particularly important to keep in mind item (b). Since science deals with the objective properties of the physical and biological world, it is a basic axiom that these properties are independent of the scientist who reported them. Everyone should be able to measure or observe the same properties within error, if they do the same experiment using the same materials and procedures. In science, one does the same experiment by exactly repeating the experiment that has been described in the Methods section. Therefore, someone can only repeat an experiment accurately if all the relevant details of the experimental methods are clearly described. 1 , 3

The following information is important to include under illustrative headings, and is generally presented in narrative form. A detailed list of all the materials used in the experiments and, if important, their source should be described. These include biological agents (e.g., zebrafish, brine shrimp), chemicals and their concentrations (e.g., 0.20 mg/mL nicotine), and physical equipment (e.g., four 10-gallon aquariums, one light timer, one 10-well falcon dish). The reader needs to know as much as necessary about each of the materials; however, it is important not to include extraneous information. For example, consider an experiment involving zebrafish. The type and characteristics of the zebrafish used must be clearly described so another scientist could accurately replicate the experiment, such as 4–6-month-old male and female zebrafish, the type of zebrafish used (e.g., Golden), and where they were obtained (e.g., the NIEHS Children's Environmental Health Sciences Core Center in the WATER Institute of the University of Wisconsin—Milwaukee). In addition to describing the physical set-up of the experiment, it may be helpful to include photographs or diagrams in the report to further illustrate the experimental design.

A thorough description of each procedure done in the reported experiment, and justification as to why a particular method was chosen to most effectively answer the research question should also be included. For example, if the scientist was using zebrafish to study developmental effects of nicotine, the reader needs to know details about how and when the zebrafish were exposed to the nicotine (e.g., maternal exposure, embryo injection of nicotine, exposure of developing embryo to nicotine in the water for a particular length of time during development), duration of the exposure (e.g., a certain concentration for 10 minutes at the two-cell stage, then the embryos were washed), how many were exposed, and why that method was chosen. The reader would also need to know the concentrations to which the zebrafish were exposed, how the scientist observed the effects of the chemical exposure (e.g., microscopic changes in structure, changes in swimming behavior), relevant safety and toxicity concerns, how outcomes were measured, and how the scientist determined whether the data/results were significantly different in experimental and unexposed control animals (statistical methods).

Students must take great care and effort to write a good Methods section because it is an essential component of the effective communication of scientific findings.

The Results section describes in detail the actual experiments that were undertaken in a clear and well-organized narrative. The information found in the Methods section serves as background for understanding these descriptions and does not need to be repeated. For each different experiment, the author may wish to provide a subtitle and, in addition, one or more introductory sentences that explains the reason for doing the experiment. In a sense, this information is an extension of the Introduction in that it makes the argument to the reader why it is important to do the experiment. The Introduction is more general; this text is more specific.

Once the reader understands the focus of the experiment, the writer should restate the hypothesis to be tested or the information sought in the experiment. For example, “Atrazine is routinely used as a crop pesticide. It is important to understand whether it affects organisms that are normally found in soil. We decided to use worms as a test organism because they are important members of the soil community. Because atrazine damages nerve cells, we hypothesized that exposure to atrazine will inhibit the ability of worms to do locomotor activities. In the first experiment, we tested the effect of the chemical on burrowing action.”

Then, the experiments to be done are described and the results entered. In reporting on experimental design, it is important to identify the dependent and independent variables clearly, as well as the controls. The results must be shown in a way that can be reproduced by the reader, but do not include more details than needed for an effective analysis. Generally, meaningful and significant data are gathered together into tables and figures that summarize relevant information, and appropriate statistical analyses are completed based on the data gathered. Besides presenting each of these data sources, the author also provides a written narrative of the contents of the figures and tables, as well as an analysis of the statistical significance. In the narrative, the writer also connects the results to the aims of the experiment as described above. Did the results support the initial hypothesis? Do they provide the information that was sought? Were there problems in the experiment that compromised the results? Be careful not to include an interpretation of the results; that is reserved for the Discussion section.

The writer then moves on to the next experiment. Again, the first paragraph is developed as above, except this experiment is seen in the context of the first experiment. In other words, a story is being developed. So, one commonly refers to the results of the first experiment as part of the basis for undertaking the second experiment. “In the first experiment we observed that atrazine altered burrowing activity. In order to understand how that might occur, we decided to study its impact on the basic biology of locomotion. Our hypothesis was that atrazine affected neuromuscular junctions. So, we did the following experiment..”

The Results section includes a focused critical analysis of each experiment undertaken. A hallmark of the scientist is a deep skepticism about results and conclusions. “Convince me! And then convince me again with even better experiments.” That is the constant challenge. Without this basic attitude of doubt and willingness to criticize one's own work, scientists do not get to the level of concern about experimental methods and results that is needed to ensure that the best experiments are being done and the most reproducible results are being acquired. Thus, it is important for students to state any limitations or weaknesses in their research approach and explain assumptions made upfront in this section so the validity of the research can be assessed.

The Discussion section is the where the author takes an overall view of the work presented in the article. First, the main results from the various experiments are gathered in one place to highlight the significant results so the reader can see how they fit together and successfully test the original hypotheses of the experiment. Logical connections and trends in the data are presented, as are discussions of error and other possible explanations for the findings, including an analysis of whether the experimental design was adequate. Remember, results should not be restated in the Discussion section, except insofar as it is absolutely necessary to make a point.

Second, the task is to help the reader link the present work with the larger body of knowledge that was portrayed in the Introduction . How do the results advance the field, and what are the implications? What does the research results mean? What is the relevance? 1 , 3

Lastly, the author may suggest further work that needs to be done based on the new knowledge gained from the research.

Supporting Documentation and Writing Skills

Tables and figures are included to support the content of the research paper. These provide the reader with a graphic display of information presented. Tables and figures must have illustrative and descriptive titles, legends, interval markers, and axis labels, as appropriate; should be numbered in the order that they appear in the report; and include explanations of any unusual abbreviations.

The final section of the scientific article is the Reference section. When citing sources, it is important to follow an accepted standardized format, such as CSE (Council of Science Editors), APA (American Psychological Association), MLA (Modern Language Association), or CMS (Chicago Manual of Style). References should be listed in alphabetical order and original authors cited. All sources cited in the text must be included in the Reference section. 1

When writing a scientific paper, the importance of writing concisely and accurately to clearly communicate the message should be emphasized to students. 1 – 3 Students should avoid slang and repetition, as well as abbreviations that may not be well known. 1 If an abbreviation must be used, identify the word with the abbreviation in parentheses the first time the term is used. Using appropriate and correct grammar and spelling throughout are essential elements of a well-written report. 1 , 3 Finally, when the article has been organized and formatted properly, students are encouraged to peer review to obtain constructive criticism and then to revise the manuscript appropriately. Good scientific writing, like any kind of writing, is a process that requires careful editing and revision. 1

A key dimension of NRC's A Framework for K–12 Science Education , Scientific and Engineering Practices, and the developing Next Generation Science Standards emphasizes the importance of students being able to ask questions, define problems, design experiments, analyze and interpret data, draw conclusions, and communicate results. 5 , 6 In the Science Education Partnership Award (SEPA) program at the University of Wisconsin—Milwaukee, we found the guidelines presented in this article useful for high school science students because this group of students (and probably most undergraduates) often lack in understanding of, and skills to develop and write, the various components of an effective scientific paper. Students routinely need to focus more on the data collected and analyze what the results indicated in relation to the research question/hypothesis, as well as develop a detailed discussion of what they learned. Consequently, teaching students how to effectively organize and write a research report is a critical component when engaging students in scientific inquiry.

Acknowledgments

This article was supported by a Science Education Partnership Award (SEPA) grant (Award Number R25RR026299) from the National Institute of Environmental Health Sciences of the National Institutes of Health. The SEPA program at the University of Wisconsin—Milwaukee is part of the Children's Environmental Health Sciences Core Center, Community Outreach and Education Core, funded by the National Institute of Environmental Health Sciences (Award Number P30ES004184). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health or the National Institute of Environmental Health Sciences.

A guide to writing scientific papers

Ids item types, copyright holder, usage metrics.

University of Zimbabwe Social Sciences Research

  • Skip to primary navigation
  • Skip to main content
  • Skip to footer

how to write scientific research report

Understanding Science

How science REALLY works...

scientific journal

Publication that contains firsthand reports of scientific research, often reviewed by experts. In these articles, scientists describe a study and any details one might need to evaluate that study — background information, data, statistical results, graphs, maps, explanations of how the study was performed and how the researchers interpreted their results, etc. To learn more, visit Publish or perish .

Subscribe to our newsletter

  • Understanding Science 101
  • The science flowchart
  • Science stories
  • Grade-level teaching guides
  • Teaching resource database
  • Journaling tool
  • Misconceptions

An image of multiple 3D shapes representing speech bubbles in a sequence, with broken up fragments of text within them.

A new ‘AI scientist’ can write science papers without any human input. Here’s why that’s a problem

how to write scientific research report

Dean, School of Computing Technologies, RMIT University, RMIT University

Disclosure statement

Karin Verspoor receives funding from the Australian Research Council, the Medical Research Future Fund, the National Health and Medical Research Council, and Elsevier BV. She is affiliated with BioGrid Australia and is a co-founder of the Australian Alliance for Artificial Intelligence in Healthcare.

RMIT University provides funding as a strategic partner of The Conversation AU.

View all partners

Scientific discovery is one of the most sophisticated human activities. First, scientists must understand the existing knowledge and identify a significant gap. Next, they must formulate a research question and design and conduct an experiment in pursuit of an answer. Then, they must analyse and interpret the results of the experiment, which may raise yet another research question.

Can a process this complex be automated? Last week, Sakana AI Labs announced the creation of an “AI scientist” – an artificial intelligence system they claim can make scientific discoveries in the area of machine learning in a fully automated way.

Using generative large language models (LLMs) like those behind ChatGPT and other AI chatbots, the system can brainstorm, select a promising idea, code new algorithms, plot results, and write a paper summarising the experiment and its findings, complete with references. Sakana claims the AI tool can undertake the complete lifecycle of a scientific experiment at a cost of just US$15 per paper – less than the cost of a scientist’s lunch.

These are some big claims. Do they stack up? And even if they do, would an army of AI scientists churning out research papers with inhuman speed really be good news for science?

How a computer can ‘do science’

A lot of science is done in the open, and almost all scientific knowledge has been written down somewhere (or we wouldn’t have a way to “know” it). Millions of scientific papers are freely available online in repositories such as arXiv and PubMed .

LLMs trained with this data capture the language of science and its patterns. It is therefore perhaps not at all surprising that a generative LLM can produce something that looks like a good scientific paper – it has ingested many examples that it can copy.

What is less clear is whether an AI system can produce an interesting scientific paper. Crucially, good science requires novelty.

But is it interesting?

Scientists don’t want to be told about things that are already known. Rather, they want to learn new things, especially new things that are significantly different from what is already known. This requires judgement about the scope and value of a contribution.

The Sakana system tries to address interestingness in two ways. First, it “scores” new paper ideas for similarity to existing research (indexed in the Semantic Scholar repository). Anything too similar is discarded.

Second, Sakana’s system introduces a “peer review” step – using another LLM to judge the quality and novelty of the generated paper. Here again, there are plenty of examples of peer review online on sites such as openreview.net that can guide how to critique a paper. LLMs have ingested these, too.

AI may be a poor judge of AI output

Feedback is mixed on Sakana AI’s output. Some have described it as producing “ endless scientific slop ”.

Even the system’s own review of its outputs judges the papers weak at best. This is likely to improve as the technology evolves, but the question of whether automated scientific papers are valuable remains.

The ability of LLMs to judge the quality of research is also an open question. My own work (soon to be published in Research Synthesis Methods ) shows LLMs are not great at judging the risk of bias in medical research studies, though this too may improve over time.

Sakana’s system automates discoveries in computational research, which is much easier than in other types of science that require physical experiments. Sakana’s experiments are done with code, which is also structured text that LLMs can be trained to generate.

AI tools to support scientists, not replace them

AI researchers have been developing systems to support science for decades. Given the huge volumes of published research, even finding publications relevant to a specific scientific question can be challenging.

Specialised search tools make use of AI to help scientists find and synthesise existing work. These include the above-mentioned Semantic Scholar, but also newer systems such as Elicit , Research Rabbit , scite and Consensus .

Text mining tools such as PubTator dig deeper into papers to identify key points of focus, such as specific genetic mutations and diseases, and their established relationships. This is especially useful for curating and organising scientific information.

Machine learning has also been used to support the synthesis and analysis of medical evidence, in tools such as Robot Reviewer . Summaries that compare and contrast claims in papers from Scholarcy help to perform literature reviews.

All these tools aim to help scientists do their jobs more effectively, not to replace them.

AI research may exacerbate existing problems

While Sakana AI states it doesn’t see the role of human scientists diminishing, the company’s vision of “a fully AI-driven scientific ecosystem” would have major implications for science.

One concern is that, if AI-generated papers flood the scientific literature, future AI systems may be trained on AI output and undergo model collapse . This means they may become increasingly ineffectual at innovating.

However, the implications for science go well beyond impacts on AI science systems themselves.

There are already bad actors in science, including “paper mills” churning out fake papers . This problem will only get worse when a scientific paper can be produced with US$15 and a vague initial prompt.

The need to check for errors in a mountain of automatically generated research could rapidly overwhelm the capacity of actual scientists. The peer review system is arguably already broken , and dumping more research of questionable quality into the system won’t fix it.

Science is fundamentally based on trust. Scientists emphasise the integrity of the scientific process so we can be confident our understanding of the world (and now, the world’s machines) is valid and improving.

A scientific ecosystem where AI systems are key players raises fundamental questions about the meaning and value of this process, and what level of trust we should have in AI scientists. Is this the kind of scientific ecosystem we want?

  • Artificial intelligence (AI)
  • Computer science
  • Research integrity
  • Paper mills

how to write scientific research report

Project Manager SSTP

how to write scientific research report

Head of Evidence to Action

how to write scientific research report

Supply Chain - Assistant/Associate Professor (Tenure-Track)

how to write scientific research report

OzGrav Postdoctoral Research Fellow

how to write scientific research report

Casual Facilitator: GERRIC Student Programs - Arts, Design and Architecture

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

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 19 August 2024

Chronic adaptive deep brain stimulation versus conventional stimulation in Parkinson’s disease: a blinded randomized feasibility trial

  • Carina R. Oehrn   ORCID: orcid.org/0000-0001-8451-7960 1   na1 ,
  • Stephanie Cernera 1   na1 ,
  • Lauren H. Hammer 2   na1 ,
  • Maria Shcherbakova 1 ,
  • Jiaang Yao   ORCID: orcid.org/0000-0001-7062-2508 1 , 3 ,
  • Amelia Hahn 1 ,
  • Sarah Wang 2 , 4 ,
  • Jill L. Ostrem 2 , 4 ,
  • Simon Little   ORCID: orcid.org/0000-0001-6249-6230 2 , 3 , 4   na2 &
  • Philip A. Starr   ORCID: orcid.org/0000-0003-2733-4003 1 , 3 , 4   na2  

Nature Medicine ( 2024 ) Cite this article

3091 Accesses

854 Altmetric

Metrics details

  • Biomedical engineering
  • Neurophysiology
  • Parkinson's disease

Deep brain stimulation (DBS) is a widely used therapy for Parkinson’s disease (PD) but lacks dynamic responsiveness to changing clinical and neural states. Feedback control might improve therapeutic effectiveness, but the optimal control strategy and additional benefits of ‘adaptive’ neurostimulation are unclear. Here we present the results of a blinded randomized cross-over pilot trial aimed at determining the neural correlates of specific motor signs in individuals with PD and the feasibility of using these signals to drive adaptive DBS. Four male patients with PD were recruited from a population undergoing DBS implantation for motor fluctuations, with each patient receiving adaptive DBS and continuous DBS. We identified stimulation-entrained gamma oscillations in the subthalamic nucleus or motor cortex as optimal markers of high versus low dopaminergic states and their associated residual motor signs in all four patients. We then demonstrated improved motor symptoms and quality of life with adaptive compared to clinically optimized standard stimulation. The results of this pilot trial highlight the promise of personalized adaptive neurostimulation in PD based on data-driven selection of neural signals. Furthermore, these findings provide the foundation for further larger clinical trials to evaluate the efficacy of personalized adaptive neurostimulation in PD and other neurological disorders. ClinicalTrials.gov registration: NCT03582891 .

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

24,99 € / 30 days

cancel any time

Subscribe to this journal

Receive 12 print issues and online access

195,33 € per year

only 16,28 € per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

how to write scientific research report

Similar content being viewed by others

how to write scientific research report

Emerging technologies for improved deep brain stimulation

how to write scientific research report

Eight-hours conventional versus adaptive deep brain stimulation of the subthalamic nucleus in Parkinson’s disease

how to write scientific research report

Modulation of subthalamic beta oscillations by movement, dopamine, and deep brain stimulation in Parkinson’s disease

Data availability.

De-identified individual participant data, including neural, wearable and digital diary data, are shared on the Data Archive for the BRAIN Initiative website ( https://dabi.loni.usc.edu/ ; https://doi.org/10.18120/cq9c-d057 ). The study protocol is provided in the Supplementary Information . The Food and Drug Administration investigational device exemption is available on the Open Mind website ( https://osf.io/cmndq/ ). Data will be available permanently with no restrictions, for purposes of replicating the findings or conducting meta-analyses.

Code availability

Code written in C# and MATLAB, which operates the investigational device and extracts raw neural data, is available on the Open Mind GitHub platform ( https://openmind-consortium.github.io ). The code for biomarker identification implemented in MATLAB is available in the repository Code Ocean, without restrictions 59 , except for code related to linear discriminant analysis (Fig. 4c–e ), which will be made available after publication of a subsequent manuscript (currently in preparation) that uses this code.

Lozano, A. M. et al. Deep brain stimulation: current challenges and future directions. Nat. Rev. Neurol. 15 , 148–160 (2019).

Article   PubMed   PubMed Central   Google Scholar  

Neumann, W. -J., Gilron, R., Little, S. & Tinkhauser, G. Adaptive deep brain stimulation: from experimental evidence toward practical implementation. Mov. Disord . https://doi.org/10.1002/mds.29415 (2023).

Marceglia, S. et al. Deep brain stimulation: is it time to change gears by closing the loop? J. Neural Eng. 18 , 061001 (2021).

Article   Google Scholar  

Stanslaski, S. et al. Design and validation of a fully implantable, chronic, closed-loop neuromodulation device with concurrent sensing and stimulation. IEEE Trans. Neural Syst. Rehabil. Eng. 20 , 410–421 (2012).

Article   PubMed   Google Scholar  

Stanslaski, S. et al. A chronically implantable neural coprocessor for investigating the treatment of neurological disorders. IEEE Trans. Biomed. Circuits Syst. 12 , 1230–1245 (2018).

Thenaisie, Y. et al. Towards adaptive deep brain stimulation: clinical and technical notes on a novel commercial device for chronic brain sensing. J. Neural Eng. 18 , 042002 (2021).

Starr, P. A. Totally implantable bidirectional neural prostheses: a flexible platform for innovation in neuromodulation. Front. Neurosci. 12 , 619 (2018).

Nakajima, A. et al. Case report: chronic adaptive deep brain stimulation personalizing therapy based on Parkinsonian state. Front. Hum. Neurosci. 15 , 702961 (2021).

Gilron, R. et al. Long-term wireless streaming of neural recordings for circuit discovery and adaptive stimulation in individuals with Parkinson’s disease. Nat. Biotechnol. 39 , 1078–1085 (2021).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Little, S. & Brown, P. Debugging adaptive deep brain stimulation for Parkinson’s disease. Mov. Disord. 35 , 555–561 (2020).

Wilkins, K. B., Melbourne, J. A., Akella, P. & Bronte-Stewart, H. M. Unraveling the complexities of programming neural adaptive deep brain stimulation in Parkinson’s disease. Front. Hum. Neurosci. 17 , 1310393 (2023).

Ansó, J. et al. Concurrent stimulation and sensing in bi-directional brain interfaces: a multi-site translational experience. J. Neural Eng. 19 , 026025 (2022).

Ascherio, A. & Schwarzschild, M. A. The epidemiology of Parkinson’s disease: risk factors and prevention. Lancet Neurol. 15 , 1257–1272 (2016).

Vitek, J. L. et al. Subthalamic nucleus deep brain stimulation with a multiple independent constant current-controlled device in Parkinson’s disease (INTREPID): a multicentre, double-blind, randomised, sham-controlled study. Lancet Neurol. 19 , 491–501 (2020).

Article   CAS   PubMed   Google Scholar  

Okun, M. S. et al. Subthalamic deep brain stimulation with a constant-current device in Parkinson’s disease: an open-label randomised controlled trial. Lancet Neurol. 11 , 140–149 (2012).

Weaver, F. M. et al. Bilateral deep brain stimulation vs best medical therapy for patients with advanced Parkinson disease: a randomized controlled trial. JAMA 301 , 63–73 (2009).

Deuschl, G. et al. A randomized trial of deep-brain stimulation for Parkinson’s disease. N. Engl. J. Med. 355 , 896–908 (2006).

Follett, K. A. et al. Pallidal versus subthalamic deep-brain stimulation for Parkinson’s disease. N. Engl. J. Med. 362 , 2077–2091 (2010).

Odekerken, V. J. et al. Subthalamic nucleus versus globus pallidus bilateral deep brain stimulation for advanced Parkinson’s disease (NSTAPS study): a randomised controlled trial. Lancet Neurol. 12 , 37–44 (2013).

Bronte-Stewart, H. et al. Adaptive DBS Algorithm for Personalized Therapy in Parkinson’s Disease: ADAPT-PD clinical trial methodology and early data (P1-11.002). Neurology https://doi.org/10.1212/WNL.0000000000203099 (2023).

Marceglia, S. et al. Double-blind cross-over pilot trial protocol to evaluate the safety and preliminary efficacy of long-term adaptive deep brain stimulation in patients with Parkinson’s disease. BMJ Open 12 , e049955 (2022).

Kühn, A. A., Kupsch, A., Schneider, G.-H. & Brown, P. Reduction in subthalamic 8-35 Hz oscillatory activity correlates with clinical improvement in Parkinson’s disease. Eur. J. Neurosci. 23 , 1956–1960 (2006).

Kühn, A. A. et al. High-frequency stimulation of the subthalamic nucleus suppresses oscillatory β activity in patients with Parkinson’s disease in parallel with improvement in motor performance. J. Neurosci. 28 , 6165–6173 (2008).

Little, S. et al. Adaptive deep brain stimulation in advanced Parkinson disease. Ann. Neurol. 74 , 449–457 (2013).

Velisar, A. et al. Dual threshold neural closed loop deep brain stimulation in Parkinson disease patients. Brain Stimul. 12 , 868–876 (2019).

Bocci, T. et al. Eight-hours conventional versus adaptive deep brain stimulation of the subthalamic nucleus in Parkinson’s disease. NPJ Park. Dis. 7 , 88 (2021).

Article   CAS   Google Scholar  

Tinkhauser, G. et al. The modulatory effect of adaptive deep brain stimulation on beta bursts in Parkinson’s disease. Brain J. Neurol. 140 , 1053–1067 (2017).

Bronstein, J. M. et al. Deep brain stimulation for Parkinson disease: an expert consensus and review of key issues. Arch. Neurol. 68 , 165 (2011).

Swann, N. C. et al. Gamma oscillations in the hyperkinetic state detected with chronic human brain recordings in Parkinson’s disease. J. Neurosci. 36 , 6445–6458 (2016).

Swann, N. C. et al. Adaptive deep brain stimulation for Parkinson’s disease using motor cortex sensing. J. Neural Eng. 15 , 046006 (2018).

Bove, F., Genovese, D. & Moro, E. Developments in the mechanistic understanding and clinical application of deep brain stimulation for Parkinson’s disease. Expert Rev. Neurother. 22 , 789–803 (2022).

Wiest, C. et al. Finely-tuned gamma oscillations: spectral characteristics and links to dyskinesia. Exp. Neurol. 351 , 113999 (2022).

Sermon, J. J. et al. Sub-harmonic entrainment of cortical gamma oscillations to deep brain stimulation in Parkinson’s disease: model based predictions and validation in three human subjects. Brain Stimul. 16 , 1412–1424 (2023).

Olaru, M. et al. Motor network gamma oscillations in chronic home recordings predict dyskinesia in Parkinson’s disease. Brain J. Neurol . https://doi.org/10.1093/brain/awae004 (2024).

Herdman, M. et al. Development and preliminary testing of the new five-level version of EQ-5D (EQ-5D-5L). Qual. Life Res. 20 , 1727–1736 (2011).

Horne, M. K., McGregor, S. & Bergquist, F. An objective fluctuation score for Parkinson’s disease. PLoS ONE 10 , e0124522 (2015).

Nutt, J. G., Woodward, W. R., Hammerstad, J. P., Carter, J. H. & Anderson, J. L. The “on–off” phenomenon in Parkinson’s disease: relation to levodopa absorption and transport. N. Engl. J. Med. 310 , 483–488 (1984).

van Rheede, J. J. et al. Diurnal modulation of subthalamic beta oscillatory power in Parkinson’s disease patients during deep brain stimulation. NPJ Parkinsons Dis. 8 , 88 (2022).

Tinkhauser, G. & Moraud, E. M. Controlling clinical states governed by different temporal dynamics with closed-loop deep brain stimulation: a principled framework. Front. Neurosci. 15 , 734186 (2021).

Alagapan, S. et al. Cingulate dynamics track depression recovery with deep brain stimulation. Nature 622 , 130–138 (2023).

Heck, C. N. et al. Two-year seizure reduction in adults with medically intractable partial onset epilepsy treated with responsive neurostimulation: final results of the RNS System Pivotal trial. Epilepsia 55 , 432–441 (2014).

Scangos, K. W. et al. Closed-loop neuromodulation in an individual with treatment-resistant depression. Nat. Med. 27 , 1696–1700 (2021).

Vizcarra, J. A. et al. Subthalamic deep brain stimulation and levodopa in Parkinson’s disease: a meta-analysis of combined effects. J. Neurol. 266 , 289–297 (2019).

Brown, P. et al. Dopamine dependency of oscillations between subthalamic nucleus and pallidum in Parkinson’s disease. J. Neurosci. 21 , 1033–1038 (2001).

Halje, P. et al. Levodopa-induced dyskinesia is strongly associated with resonant cortical oscillations. J. Neurosci. 32 , 16541–16551 (2012).

Wiest, C. et al. Subthalamic deep brain stimulation induces finely-tuned gamma oscillations in the absence of levodopa. Neurobiol. Dis. 152 , 105287 (2021).

Arlotti, M. et al. Eight-hours adaptive deep brain stimulation in patients with Parkinson disease. Neurology 90 , e971–e976 (2018).

Foffani, G. & Alegre, M. Brain oscillations and Parkinson disease. Handb. Clin. Neurol. 184 , 259–271 (2022).

Feldmann, L. K. et al. Toward therapeutic electrophysiology: beta-band suppression as a biomarker in chronic local field potential recordings. NPJ Parkinsons Dis. 8 , 44 (2022).

Chen, Y. et al. Neuromodulation effects of deep brain stimulation on beta rhythm: a longitudinal local field potential study. Brain Stimul. 13 , 1784–1792 (2020).

Olson, J. D. et al. Comparison of subdural and subgaleal recordings of cortical high-gamma activity in humans. Clin. Neurophysiol. 127 , 277–284 (2016).

Piña-Fuentes, D. et al. Acute effects of adaptive deep brain stimulation in Parkinson’s disease. Brain Stimul. 13 , 1507–1516 (2020).

Busch, J. L. et al. Single threshold adaptive deep brain stimulation in Parkinson’s disease depends on parameter selection, movement state and controllability of subthalamic beta activity. Brain Stimul. 17 , 125–133 (2024).

Merk, T. et al. Machine learning based brain signal decoding for intelligent adaptive deep brain stimulation. Exp. Neurol. 351 , 113993 (2022).

Davis, T. S. et al. LeGUI: a fast and accurate graphical user interface for automated detection and anatomical localization of intracranial electrodes. Front. Neurosci. 15 , 769872 (2021).

Horn, A. et al. Lead-DBS v2: towards a comprehensive pipeline for deep brain stimulation imaging. NeuroImage 184 , 293–316 (2019).

Oehrn, C. R. et al. Chronic adaptive deep brain stimulation is superior to conventional stimulation in Parkinson's disease: a blinded randomized feasibility trial [Source Data]. Data Archive for the Brain Initiative https://doi.org/10.18120/cq9c-d057 (2024).

Oostenveld, R., Fries, P., Maris, E. & Schoffelen, J. -M. FieldTrip: open source software for advanced analysis of MEG, EEG, and invasive electrophysiological data. Comput. Intell. Neurosci. 2011 , 156869 (2011).

Oehrn, C. R. et al. Chronic adaptive deep brain stimulation is superior to conventional stimulation in Parkinson's disease: a blinded randomized feasibility trial. Code Ocean . https://doi.org/10.24433/CO.5656158.v1 (2024).

Oehrn, C. R. et al. Direct electrophysiological evidence for prefrontal control of hippocampal processing during voluntary forgetting. Curr. Biol. 28 , 3016–3022 (2018).

Maris, E. & Oostenveld, R. Nonparametric statistical testing of EEG- and MEG-data. J. Neurosci. Methods 164 , 177–190 (2007).

Gilron, R. et al. Sleep-aware adaptive deep brain stimulation control: chronic use at home with dual independent linear discriminate detectors. Front. Neurosci. 15 , 732499 (2021).

Cernera, S. et al. Wearable sensor-driven responsive deep brain stimulation for essential tremor. Brain Stimul. 14 , 1434–1443 (2021).

Hammer, L. H., Kochanski, R. B., Starr, P. A. & Little, S. Artifact characterization and a multipurpose template-based offline removal solution for a sensing-enabled deep brain stimulation device. Stereotact. Funct. Neurosurg. 100 , 168–183 (2022).

Neumann, W. -J. et al. The sensitivity of ECG contamination to surgical implantation site in brain computer interfaces. Brain Stimul. 14 , 1301–1306 (2021).

Goetz, C. G. et al. Movement Disorder Society-sponsored revision of the Unified Parkinson’s Disease Rating Scale (MDS-UPDRS): process, format, and clinimetric testing plan. Mov. Disord. 22 , 41–47 (2007).

McAuley, M. D. Incorrect calculation of total electrical energy delivered by a deep brain stimulator. Brain Stimul. 13 , 1414–1415 (2020).

Download references

Acknowledgements

The study was supported by National Institute of Neurological Disorders and Stroke (NINDS) UH3NS100544 (to P.A.S.), the Parkinson Fellowship of the Thiemann Foundation (to C.R.O.), NINDS F32NS129627 (to S.C.), NINDS R25NS070680 (to L.H.H.) and TUYF Charitable Trust Fund (to J.Y.). Research reported in this publication was also partly supported by R01 NS090913 (to P.A.S.), NINDS K23NS120037 (to S.L.) and R01 NS131405 (to S.L.). Investigational devices were provided at no charge by the manufacturer, but the manufacturer had no role in the conduct, analysis or interpretation of the study. The Open Mind consortium for technology dissemination, funded by NINDS U24 NS113637 (to P.A.S.), provided technical resources for the use of the Summit RC+S neural interface. We thank T. Wozny for lead localization, W. Chiong for neuroethical input, C. Smyth, R. Gilron, R. Wilt and C. de Hemptinne for technical contributions and K. Probst for medical art (Fig. 1a ). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Author information

These authors contributed equally: Carina R. Oehrn, Stephanie Cernera, Lauren H. Hammer.

These authors jointly supervised this work: Simon Little, Philip A Starr.

Authors and Affiliations

Department of Neurological Surgery, University of California, San Francisco, San Francisco, CA, USA

Carina R. Oehrn, Stephanie Cernera, Maria Shcherbakova, Jiaang Yao, Amelia Hahn & Philip A. Starr

Department of Neurology, University of California, San Francisco, San Francisco, CA, USA

Lauren H. Hammer, Sarah Wang, Jill L. Ostrem & Simon Little

Graduate Program in Bioengineering, University of California, Berkeley and University of California, San Francisco, San Francisco, CA, USA

Jiaang Yao, Simon Little & Philip A. Starr

Weill Institute for Neurosciences, University of California, San Francisco, San Francisco, CA, USA

Sarah Wang, Jill L. Ostrem, Simon Little & Philip A. Starr

You can also search for this author in PubMed   Google Scholar

Contributions

P.A.S., S.L., J.L.O., C.R.O., S.C. and L.H.H. designed the study and analysis pipeline. C.R.O., S.C., L.H.H., M.S. and J.Y. collected and analyzed the data. A.H. facilitated patient communication and coordination throughout the study. S.W. oversaw study administration, including institutional review board approval and regulatory compliance. C.R.O., S.C., L.H.H., S.L. and P.A.S. drafted the manuscript, and all authors reviewed, commented on and approved the final version.

Corresponding author

Correspondence to Carina R. Oehrn .

Ethics declarations

Competing interests.

S.L. consults for Iota Biosciences. J.L.O. reports support from Medtronic and Boston Scientific for research and education and consults for AbbVie and Rune Labs. P.A.S. receives support from Medtronic and Boston Scientific for fellowship education. C.R.O., S.C., L.H.H., M.S., J.Y., A.H. and S.W. declare no competing interests.

Peer review

Peer review information.

Nature Medicine thanks Jaimie Henderson, Andrea Kühn and Theoden Netoff for their contribution to the peer review of this work. Primary Handling Editor: Jerome Staal, in collaboration with the Nature Medicine team.

Additional information

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

Extended data

Extended data fig. 1 localization of leads over sensorimotor cortex and within subthalamic nucleus in native space..

a–d , Example localization of cortical and subcortical leads in patient 2, generated by fusing postoperative CT with preoperative MRI scans. Contacts appear as white CT artifacts due to metal content and are labeled with red arrows. a , Cortical leads on axial T1-weighted MRI through the vertex. b , STN leads on axial T2-weighted MRI through the region of the dorsal STN, 3 mm inferior to the intercommissural plane. c,d , Cortical leads on oblique sagittal T1-weighted MRI passing through the long axis of the lead array in left (c) and right (d) hemispheres, respectively. e–h , Location of cortical leads for each patient overlayed on 3D reconstruction of cortex rendered using the Locate Electrodes Graphical User Interface (LeGUI). Electrodes used in the anterior and posterior cortical montages are shown in cyan and yellow, respectively. For patient 1 (e) , 2 (f) and 4 (h) , anterior and posterior montages covered the pre- and postcentral gyrus, respectively. For patient 3, right side (g) , the anterior montage included one electrode on the middle frontal and one on the precentral gyrus. The posterior montage comprised one pre- and one postcentral electrode. In all figures, red arrows indicate the location of the central sulcus.

Extended Data Fig. 2 Initial and finalized adaptive stimulation parameters and example adaptive control policies.

a , Suggested initial parameters for algorithms developed for time scales of minutes to hours, as identified during steps 5 and 6 of the pipeline. An update rate of 10 s typically provided a signal to noise ratio that allowed for adequate discrimination between the presence and absence of the most bothersome symptom, and this could often be improved with a further increase in update rate. The ramp rate chosen for each patient depended on the results of step 5 (we chose an example of 1 mA/s). b , Detailed final adaptive stimulation parameters including control signals, thresholds, FFT interval, update rates, blanking periods, onset and termination duration, and ramp rates used for each patient and hemisphere. c–e , Examples of potential control policies that can be used for an adaptive algorithm, using artificial data. The upper subpanels of each subfigure illustrate an on-state biomarker (blue), as used in our study, along with thresholds (red). Lower subpanels demonstrate the adjustment of stimulation amplitude based on the relationship of the neural signal to the thresholds. c , A single threshold control policy with two stimulation amplitudes. When the biomarker is above the threshold, stimulation amplitude decreases and once below threshold, stimulation amplitude increases. d , A dual threshold control policy with three stimulation amplitudes (not used in this study), which may be applied to address three symptom states. When the neural signal is below both thresholds, the stimulation amplitude is high (for example, 4 mA). When the biomarker is between the two thresholds, stimulation adjusts to a middle amplitude (for example, 3 mA). When the biomarker exceeds the second threshold, stimulation decreases to the low amplitude (for example, 2 mA). e , A control policy utilizing a middle state as noise buffer. Stimulation is high when the control signal is below the bottom threshold and stimulation is low when the control signal is above the top threshold. When the control signal is between the two thresholds, it remains at the level of the stimulation amplitude prior to crossing the threshold (that is, no changes are made).

Extended Data Fig. 3 Neural biomarkers of medication effects identified in-clinic.

a,b , All tables show the results from our within-patient non-parametric cluster-based permutation analyses using in-clinic recordings during two medication states (off vs. on) and stimulation conditions (low vs. high stimulation amplitude). P -values were Bonferroni-corrected for multiple comparisons. Note that p < 10 −3 indicates that the cluster was found in all 1000 permutations. This means the probability of observing this effect by chance is less than 1 in 1000. a , Statistics for the largest main effect of medication, stimulation, and their interaction for each patient and hemisphere when searching the whole frequency space (2–100 Hz) across brain regions. Frequencies represent the center frequency of 1-Hz wide power spectral density bins. For all four patients (five out of six hemispheres), we found that gamma power (specifically, stimulation-entrained gamma in four hemispheres) in the STN or cortex was the best predictor of medication state (in pat-3L, there was no significant effect of medication in any frequency band in clinic, but at home symptom monitoring identified cortical stimulation-entrained gamma power as neural biomarker; Extended Data Fig. 4 ). Positive Cohen’s d values for the medication effect highlight that the neural biomarker was higher during on-medication states. Positive Cohen’s d values for the stimulation effect indicate that the neural biomarker was higher during on-stimulation states (independent of medication), which could result in undesirable self-triggering of the algorithm (threshold crossing of the neural biomarker linked to stimulation change itself, rather than true fluctuations of medication states and symptoms). Therefore, for patient 1, we excluded 63 and 67 Hz from the subsequently used control signal (positive Cohen’s d main effect of stimulation). For patients 2, 3 and 4, we did not find stimulation effects that positively modulated biomarkers and therefore were unrestricted in biomarker selection. b , When constraining the anatomic location and frequency space to STN beta oscillations (13–30 Hz), STN spectral beta power was only predictive for medication state in two hemispheres (pat-2R and pat-4) and smaller in effect size than cortical/STN stimulation-entrained gamma oscillations for all patients.

Extended Data Fig. 4 Neural biomarkers of symptoms identified at-home.

We identified predictors of the most bothersome symptom (pat-1: bradykinesia, pat-2: lower limb dystonia), or the opposite symptom that limits the therapeutic window (pat-3 and pat-4: dyskinesia). a , Heatmaps of t -values derived from stepwise linear regressions using 1 Hz power bands between 2–100 Hz in the STN (left), anterior cortical montage (middle) and posterior cortical montage (right) to predict symptoms continuously measured with upper extremity wearable monitors for patients 1, 3 and 4 (patient 2’s bothersome symptom did not involve the upper extremity). b–d , Results from the linear regression (left) and linear discriminant analysis (LDA; right). P-values were Bonferroni-corrected for multiple comparisons (289 predictors). b , Both methods provide converging evidence that stimulation-entrained gamma power centered at half the stimulation frequency (65 Hz) in the STN and cortex optimally distinguishes hypo- and hyperkinetic symptoms. c , When constraining the anatomic location and frequency space to STN beta oscillations (13–30 Hz), frequency bands identified as most predictive were less discriminative than cortical/STN stimulation-entrained gamma oscillations (LDA: AUC < 0.7). Regression models resulted in smaller magnitude coefficients, with only one hemisphere demonstrating a significant negative association with hyperkinetic symptoms (pat-3L). d , STN beta frequency bands were also poorly predictive of wearable bradykinesia scores (AUC < 0.6), again with only one hemisphere demonstrating a significant effect in the regression model (corresponding to a positive relationship with hypokinetic symptoms; pat-3L). e , Comparison of LDA results for STN and cortical gamma activity in predicting bothersome symptoms. Neural signals selected for adaptive stimulation are shaded in grey. In three out of six hemispheres (pat-2L, pat-2R, pat-4), stimulation-entrained gamma activity in the STN distinguished between hypo- and hyperkinetic symptoms. For pat-2, STN stimulation-entrained spectral gamma power was the optimal biomarker used for aDBS in both hemispheres. In pat-4, stimulation-entrained gamma activity in the STN was a strong predictor of residual motor signs but slightly underperformed compared to cortical signals. f , Visual illustration of AUC values comparing STN and cortical gamma activity in predicting bothersome symptoms. For pat-4, the predictive value of stimulation-entrained spectral gamma power was only slightly reduced compared to cortical signals.

Extended Data Fig. 5 Beta oscillations in the STN.

a , Power spectral density in the STN based on in-clinic recordings off medication and off stimulation for all six hemispheres. All but one hemisphere (pat-1) exhibited a peak in the beta frequency band (illustrated in yellow). b , Example of the suppressive effect of DBS on STN beta oscillations precluding use of beta band activity as a biomarker of medication state during active stimulation (pat-2L, all data collected during the same in-clinic recording session). Off stimulation, the spectral peak in the beta frequency range was suppressed by medication (13–21 Hz, Cohens’ d  = −1.09, p < 10 −3 ). However, this medication effect diminished during active stimulation, even at low stimulation amplitudes (1.8 mA, largest effect in the beta band: 15–18 Hz, Cohens’ d  = 0.31, p  = 0.026). Data are corrected for stimulation-induced broadband shifts.

Extended Data Fig. 6 Effects of aDBS and cDBS on most bothersome symptom severity, additional motor symptoms, and sleep quality.

a–j , Bar plots illustrating the mean (±s.e.m.) self-reported symptoms, aside from the most bothersome symptoms, across testing days. Each dot represents the rating for one testing day (blue: cDBS, red: aDBS). These ratings constituted secondary outcome measures to ensure that we are not aggravating other motor and non-motor symptoms. a,b , Patient self-reported motor symptom severity from daily questionnaires (1 = least severe, 10 = most severe). Note that patients rated symptom severity (shown here) independently of symptom duration ; bar graphs for the latter are in Fig. 5a,b . Patient 3 did not record ratings within the instructed range of 1–10 and their data are therefore not reported. a , In addition to a decrease in the amount of daily hours with the most bothersome symptom (symptom duration , shown in Fig. 5a ), patients 1, 2, and 4 also experienced a significant improvement of symptom severity (pat-1: p < 10 −5 , pat-2: p  = 0.018, pat = 4: p  = 0.003). b , No subject reported worsened severity of their opposite symptom (pat-1: p  = 0.18, pat-2: p  = 1, pat-4: p  = 0.19). c–h , Comprehensive list of the self-reported duration of motor symptoms from daily questionnaires. These bar graphs illustrate only symptoms that were not identified by the patient as the most bothersome or as the opposite symptom. For each patient’s most bothersome symptom, results are displayed in Fig. 5a and panel a of this figure; and are labeled in c–h as not applicable (n/a). None of these “other” motor symptoms were worsened by aDBS, and patient 2 demonstrated significant improvement in the percentage of waking hours with dyskinesia ( d , p = 0.044) and gait disturbance ( h , p < 10 −4 ). i,j , Self-reported sleep quality (1 = poorest sleep, 10 = best sleep) and duration from daily questionnaires. aDBS provided no change in patients’ sleep characteristics. The number of testing days for each patient and condition used for statistical tests are summarized in Fig. 6a . Asterisks illustrate results from two-sided Wilcoxon rank sum tests. P-values for all within-subject control analyses were adjusted for multiple comparisons using the false discovery rate procedure and are indicated as: *p < 0.05, **p < 0.01, ***p < 0.001.

Extended Data Fig. 7 aDBS algorithm dynamics during nighttime.

a , Percent time spent at each stimulation amplitude during the night. Each dot represents the mean values of one night of aDBS testing across high stimulation states (orange) and low stimulation states (blue) in one hemisphere. Graphs are standard box plots (center: median; box limits: upper and lower quartiles; whiskers: minima = 25th percentile-1.5 times the interquartile range, maxima = 75th percentile+1.5 times the interquartile range). Each patient spent most of the night in the high stimulation state. b , Mean (±s.e.m.) total electrical energy delivered (TEED) during aDBS and cDBS overnight, showing increased TEED during aDBS, similar to daytime analyses (stimulation main effect: β  = 27.7, p  < 10 −25 , time main effect: β  = 0.05, p  = 0.377). Individually, TEED was increased in all hemispheres during aDBS (two-sided, one-sample Wilcoxon signed rank test, pat-1: p < 10 −6 , pat-2R: p < 10 −5 , pat-2L: p < 10 −5 , pat-3R: p < 10 −6 , pat-3L: p < 10 −6 , pat-4: p < 10 −4 ). The number of testing nights for each patient and condition used for both illustrations are stated in Fig. 6a and are equivalent to the testing days. Asterisks illustrate results from two-sided one-sample Wilcoxon signed rank tests. P-values for TEED evaluations were adjusted for multiple comparisons using the false discovery rate procedure and are indicated as: *p < 0.05, **p < 0.01, ***p < 0.001.

Extended Data Fig. 8 Flowchart of biomarker identification analyses.

We identified neural biomarkers using standardized in-clinic and at-home recordings in patients’ naturalistic environments. Non-parametric cluster-based permutation analysis identified candidate spectral biomarkers from in-clinic data by assessing main effects of medication state, stimulation amplitude, and the interaction. Next, the predictability of neural biomarkers as robust aDBS control signals of symptom state was tested using at-home recordings. For patients where the most bothersome symptom was monitored by a wearable device (for example, upper extremity bradykinesia or dyskinesia), linear stepwise regression was used to take advantage of the continuous nature of the symptom measurements. The most predictive frequency bands and recording sites were selected based on t -values. If the patient’s most bothersome symptom could not be captured by wearable monitors, the patient’s motor diaries and streaming app entries instead labeled the presence of symptoms. A linear discriminant analysis (LDA) based method identified the most predictive frequency band and recording site from these discretely labeled neural signal data, as measured by the area under the receiver operating curve (AUC). We also applied the LDA-based approach to symptoms measured by wearable monitors by mapping the continuous wearable scores to discrete symptom labels using a patient-specific dichotomization. This dichotomization allowed for subsequent offline assessment of the prediction accuracy based on multiple neural biomarkers combined as shown in Fig. 4e (note for online aDBS only single power band classifiers were implemented, as multiple power band classifiers were not found to be superior).

Supplementary information

Supplementary information.

Supplementary Methods, Tables 1 and 2 and References.

Reporting Summary

Rights and permissions.

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

Reprints and permissions

About this article

Cite this article.

Oehrn, C.R., Cernera, S., Hammer, L.H. et al. Chronic adaptive deep brain stimulation versus conventional stimulation in Parkinson’s disease: a blinded randomized feasibility trial. Nat Med (2024). https://doi.org/10.1038/s41591-024-03196-z

Download citation

Received : 04 January 2024

Accepted : 15 July 2024

Published : 19 August 2024

DOI : https://doi.org/10.1038/s41591-024-03196-z

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

Quick links

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

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

how to write scientific research report

IMAGES

  1. FREE 8+ Sample Scientific Reports in PDF

    how to write scientific research report

  2. FREE 17+ Sample Research Reports in PDF

    how to write scientific research report

  3. FREE 10+ Scientific Research Report Templates in PDF

    how to write scientific research report

  4. how to write a method in science report

    how to write scientific research report

  5. Buy Lab Report writing. How to write lab reports?

    how to write scientific research report

  6. 13+ SAMPLE Scientific Research Report in PDF

    how to write scientific research report

COMMENTS

  1. How to Write a Scientific Report

    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.

  2. Scientific Reports

    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. Research paper Writing a scientific article: A step-by-step guide for

    Abstract Many young researchers find it extremely difficult to write scientific articles, and few receive specific training in the art of presenting their research work in written format. Yet, publication is often vital for career advancement, to obtain funding, to obtain academic qualifications, or for all these reasons. We describe here the basic steps to follow in writing a scientific ...

  4. Research Report

    Research Report is a written document that presents the results of a research project or study, including the research question, methodology, results, and conclusions, in a clear and objective manner.

  5. Toolkit: How to write a great paper

    Excellent science is an essential ingredient of any great research paper, but concise writing and a clear structure are also crucial.

  6. How to write a research paper

    How do you know when it's time to start working on a paper? What is your overall writing process? Ideally, by the time I start writing a paper I have a strong foundation for why I decided to research this topic, robust results from different experiments that support my idea, and a good overview of how my research advances scientific knowledge.

  7. Writing Center

    This page is your source for scientific writing & publishing essentials. Learn how to write a successful scientific research article with our free, practical guides and hands-on resources for authors looking to improve their scientific publishing skillset.

  8. How to Write a Research Paper

    How to Write a Research Paper | A Beginner's Guide A research paper is a piece of academic writing that provides analysis, interpretation, and argument based on in-depth independent research.

  9. How to Write and Publish a Research Paper for a Peer ...

    Communicating research findings is an essential step in the research process. Often, peer-reviewed journals are the forum for such communication, yet many researchers are never taught how to write a publishable scientific paper. In this article, we explain the basic structure of a scientific paper and describe the information that should be included in each section. We also identify common ...

  10. How to write a first-class paper

    How to write a first-class paper Six experts offer advice on producing a manuscript that will get published and pull in readers.

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

    Abstract Scientific writing, while an indispensable step of the scientific process, is often overlooked in undergraduate courses in favor of maximizing class time devoted to scientific concepts. However, the ability to effectively communicate research findings is crucial for success in the biological sciences. Graduate students are encouraged to publish early and often, and professional ...

  12. How to Write a Scientific Paper: Practical Guidelines

    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.

  13. Research Guides: Writing a Scientific Paper: RESULTS

    Discussion of how to understand and write different sections of a scientific paper. Discussions of how to write Abstract, Introduction, Methods, Data, and Discussion.

  14. PDF How to Write an Effective Research REport

    Abstract. This guide for writers of research reports consists of practical suggestions for writing a report that is clear, concise, readable, and understandable. It includes suggestions for terminology and notation and for writing each section of the report—introduction, method, results, and discussion. Much of the guide consists of ...

  15. How to Write Your First Research Paper

    Writing a research manuscript is an intimidating process for many novice writers in the sciences. One of the stumbling blocks is the beginning of the process and creating the first draft. This paper presents guidelines on how to initiate the writing process ...

  16. PDF A Step by Step Guide to Writing a Scientific Manuscript

    This manuscript is intended to provide step by step instruction on how to write a scientific manuscript. The purpose is to provide a cure for "writer's block," and thus enhance a successful scientific career.2 The audience for this manuscript is the junior academician who needs guidance on how to write a manuscript.

  17. How to Write a Scientific Paper

    A scientific paper is a report of research, prepared from the investigator&#8217;s experimental findings, and intended to contribute to knowledge. The publication process is predicated by the content (the data) and the requirements of the publication (which is...

  18. Library Research Guides: STEM: How To Write A Scientific Paper

    The introduction of a scientific paper discusses the problem being studied and other theory that is relevant to understanding the findings. The hypothesis of the experiment and the motivation for the research are stated in this section. Write the introduction in your own words. Try not to copy from a lab manual or other guidelines.

  19. A Guide to Writing a Scientific Paper: A Focus on High School Through

    This article presents a detailed guide for high school through graduate level instructors that leads students to write effective and well-organized scientific papers. Interesting research emerges from the ability to ask questions, define problems, design ...

  20. Write a Research Paper

    Read Writing a Research Paper for Your Science Fair Project to learn about the purpose of a research paper and how to write one. Review How to Write a Bibliography in APA and MLA styles With Examples to learn how to properly cite resources in your paper using in-text citations. Answer the following questions to check your learning:

  21. How to Think About Your Research

    In the following sections, we will focus on empirical research writing in any of the research types, but if you are completing other types of writing (e.g., a state-of-the-art review, model development article, or conference paper), the skills you acquire will help you write those pieces as well. They just won't be the emphasis here.

  22. A guide to writing scientific papers

    The method of presentation must be made to fit the nature of the subject, the purpose of the article and the interests of probable readers. However the outline which follows may be used in the presentation of a wide variety of papers and can be adapted for special requirements.An Occasional Paper on how to write a scientific research papers.

  23. scientific journal

    Publication that contains firsthand reports of scientific research, often reviewed by experts. In these articles, scientists describe a study and any details one might need to evaluate that study — background information, data, statistical results, graphs, maps, explanations of how the study was performed and how the researchers interpreted their results, etc.

  24. A new 'AI scientist' can write science papers without any human input

    There are already bad actors in science, including "paper mills" churning out fake papers. This problem will only get worse when a scientific paper can be produced with US$15 and a vague ...

  25. Early science and colossal stone engineering in Menga, a Neolithic

    Here, we examine a great Neolithic engineering feat: the Menga dolmen, Iberia's largest megalithic monument. As listed by UNESCO, the Antequera megalithic site includes two natural formations, La Peña de los Enamorados and El Torcal karstic massif, and four major megalithic monuments: Menga, Viera, El Romeral, and the one recently discovered at Piedras Blancas, at the foot of La Peña de ...

  26. Adobe Workfront

    Add, edit, and manage the types of report widgets you want to add to dashboards to quickly build datasets. Export your Workfront data into data lakes or other business intelligence (BI) tools. Learn how to measure and report. Helping some of the world's biggest companies build a marketing system of record.

  27. PDF Scientific Reports

    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 ...

  28. Chronic adaptive deep brain stimulation versus conventional ...

    J.L.O. reports support from Medtronic and Boston Scientific for research and education and consults for AbbVie and Rune Labs. P.A.S. receives support from Medtronic and Boston Scientific for ...