Sample Paper in Scientific Format

Biology 151/152.

The sample paper below has been compressed into the left-hand column on the pages below. In the right-hand column we have included notes explaining how and why the paper is written as it is.

Article type icon

How to Format a Scientific Paper

#scribendiinc

Written by  Joanna Kimmerly-Smith

You've done the research. You've carefully recorded your lab results and compiled a list of relevant sources. You've even written a draft of your scientific, technical, or medical paper, hoping to get published in a reputable journal. But how do you format your paper to ensure that every detail is correct? If you're a scientific researcher or co-author looking to get your research published, read on to find out how to format your paper.

While it's true that you'll eventually need to tailor your research for your target journal, which will provide specific author guidelines for formatting the paper (see, for example, author guidelines for publications by Elsevier , PLOS ONE , and  mBio ), there are some formatting rules that are useful to know for your initial draft. This article will explore some of the formatting rules that apply to all scientific writing, helping you to follow the correct order of sections ( IMRaD ), understand the requirements of each section, find resources for standard terminology and units of measurement, and prepare your scientific paper for publication.

Format Overview

The four main elements of a scientific paper can be represented by the acronym IMRaD: introduction, methods, results, and discussion. Other sections, along with a suggested length,* are listed in the table below.

* Length guidelines are taken from https://www.elsevier.com/connect/11-steps-to-structuring-a-science-paper-editors-will-take-seriously#step6 .

Now, let's go through the main sections you might have to prepare to format your paper.

On the first page of the paper, you must present the title of the paper along with the authors' names, institutional affiliations, and contact information. The corresponding author(s) (i.e., the one[s] who will be in contact with the reviewers) must be specified, usually with a footnote or an asterisk (*), and their full contact details (e.g., email address and phone number) must be provided. For example:

Dr. Clara A. Bell 1, * and Dr. Scott C. Smith 2

1 University of Areopagitica, Department of Biology, Sometown, Somecountry

2 Leviathan University, Department of Biochemistry and Biomedical Sciences, Sometown, Somecountry

*[email protected]

FORMATTING TIPS:

  • If you are unsure of how to classify author roles (i.e., who did what), guidelines are available online. For example, American Geophysical Union (AGU) journals now recommend using Contributor Roles Taxonomy (CRediT), an online taxonomy for author contributions.

In this summary of your research, you must state your subject (i.e., what you did) and encapsulate the main findings and conclusions of your paper.

  • Do not add citations in an abstract (the reader might not be able to access your reference list).
  • Avoid using acronyms and abbreviations in the abstract, as the reader may not be familiar with them. Use full terms instead.

Below the abstract, include a list of key terms to help other researchers locate your study. Note that "keywords" is one word (with no space) and is followed by a colon:

Keywords : paper format, scientific writing.

  • Check whether "Keywords" should be italicized and whether each term should be capitalized.
  • Check the use of punctuation (e.g., commas versus semicolons, the use of the period at the end).
  • Some journals (e.g., IEEE ) provide a taxonomy of keywords. This aids in the classification of your research.

Introduction

This is the reader's first impression of your paper, so it should be clear and concise. Include relevant background information on your topic, using in-text citations as necessary. Report new developments in the field, and state how your research fills gaps in the existing research. Focus on the specific problem you are addressing, along with its possible solutions, and outline the limitations of your study. You can also include a research question, hypothesis, and/or objectives at the end of this section.

  • Organize your information from broad to narrow (general to particular). However, don't start too broad; keep the information relevant.
  • You can use in-text citations in this section to situate your research within the body of literature.

This is the part of your paper that explains how the research was done. You should relate your research procedures in a clear, logical order (i.e., the order in which you conducted the research) so that other researchers can reproduce your results. Simply refer to the established methods you used, but describe any procedures that are original to your study in more detail.

  • Identify the specific instruments you used in your research by including the manufacturer’s name and location in parentheses.
  • Stay consistent with the order in which information is presented (e.g., quantity, temperature, stirring speed, refrigeration period).

Now that you've explained how you gathered your research, you've got to report what you actually found. In this section, outline the main findings of your research. You need not include too many details, particularly if you are using tables and figures. While writing this section, be consistent and use the smallest number of words necessary to convey your statistics.

  • Use appendices or supplementary materials if you have too much data.
  • Use headings to help the reader follow along, particularly if your data are repetitive (but check whether your style guide allows you to use them).

In this section, you interpret your findings for the reader in relation to previous research and the literature as a whole. Present your general conclusions, including an assessment of the strengths and weaknesses of the research and the implications of your findings. Resolve the hypothesis and/or research question you identified in the introduction.

  • Use in-text citations to support your discussion.
  • Do not repeat the information you presented in the results or the introduction unless it is necessary for a discussion of the overall implications of the research.

This section is sometimes included in the last paragraph of the discussion. Explain how your research fits within your field of study, and identify areas for future research.

  • Keep this section short.

Acknowledgments

Write a brief paragraph giving credit to any institution responsible for funding the study (e.g., through a fellowship or grant) and any individual(s) who contributed to the manuscript (e.g., technical advisors or editors).

  • Check whether your journal uses standard identifiers for funding agencies (e.g., Elsevier's Funder Registry ).

Conflicts of Interest/Originality Statement

Some journals require a statement attesting that your research is original and that you have no conflicts of interest (i.e., ulterior motives or ways in which you could benefit from the publication of your research). This section only needs to be a sentence or two long.

Here you list citation information for each source you used (i.e., author names, date of publication, title of paper/chapter, title of journal/book, and publisher name and location). The list of references can be in alphabetical order (author–date style of citation) or in the order in which the sources are presented in the paper (numbered citations). Follow your style guide; if no guidelines are provided, choose a citation format and be consistent .

  • While doing your final proofread, ensure that the reference list entries are consistent with the in-text citations (i.e., no missing or conflicting information).
  • Many citation styles use a hanging indent and may be alphabetized. Use the styles in Microsoft Word to aid you in citation format.
  • Use EndNote , Mendeley , Zotero , RefWorks , or another similar reference manager to create, store, and utilize bibliographic information.

Appendix/Supplementary Information

In this optional section, you can present nonessential information that further clarifies a point without burdening the body of the paper. That is, if you have too much data to fit in a (relatively) short research paper, move anything that's not essential to this section.

  • Note that this section is uncommon in published papers. Before submission, check whether your journal allows for supplementary data, and don't put any essential information in this section.

Beyond IMRaD: Formatting the Details

Aside from the overall format of your paper, there are still other details to watch out for. The sections below cover how to present your terminology, equations, tables and figures, measurements, and statistics consistently based on the conventions of scientific writing.

Terminology

Stay consistent with the terms you use. Generally, short forms can be used once the full term has been introduced:

  • full terms versus acronyms (e.g., deoxyribonucleic acid versus DNA);
  • English names versus Greek letters (e.g., alpha versus α); and
  • species names versus short forms (e.g., Staphylococcus aureus versus S. aureus ).

One way to ensure consistency is to use standard scientific terminology. You can refer to the following resources, but if you're not sure which guidelines are preferred, check with your target journal.

  • For gene classification, use GeneCards , The Mouse Genome Informatics Database , and/or genenames.org .
  • For chemical nomenclature, refer to the International Union of Pure and Applied Chemistry (IUPAC) Compendium of Chemical Terminology (the Gold Book ) and the  IUPAC–IUB Combined Commission on Biochemical Nomenclature .
  • For marine species names, use the World Register of Marine Species (WoRMS) or the European Register of Marine Species (ERMS) .

Italics must be used correctly for scientific terminology. Here are a couple of formatting tips:

  • Species names, which are usually in Greek or Latin, are italicized (e.g., Staphylococcus aureus ).
  • Genes are italicized, but proteins aren't.

Whether in mathematical, scientific, or technical papers, equations follow a conventional format. Here are some tips for formatting your calculations:

  • Number each equation you present in the text, inserting the number in parentheses.

X + Y = 1                                                                                                                                               (1)

  • Check whether your target journal requires you to capitalize the word "Equation" or use parentheses for the equation number when you refer to equations within the text.

In Equation 1, X represents . . .

In equation (1), X represents . . .

(Note also that you should use italics for variables.)

  • Try using MathType or Equation Editor in Microsoft Word to type your equations, but use Unicode characters when typing single variables or mathematical operators (e.g., x, ≥, or ±) in running text. This makes it easier to edit your text and format your equations before publication.
  • In line with the above tip, remember to save your math equations as editable text and not as images in case changes need to be made before publication.

Tables and Figures

Do you have any tables, graphs, or images in your research? If so, you should become familiar with the rules for referring to tables and figures in your scientific paper. Some examples are presented below.

  • Capitalize the titles of specific tables and figures when you refer to them in the text (e.g., "see Table 3"; "in Figure 4").
  • In tables, stay consistent with the use of title case (i.e., Capitalizing Each Word) and sentence case (i.e., Capitalizing the first word).
  • In figure captions, stay consistent with the use of punctuation, italics, and capitalization. For example:

Figure 1. Classification of author roles.

Figure 2: taxonomy of paper keywords

Measurements

Although every journal has slightly different formatting guidelines, most agree that the gold standard for units of measurement is the International System of Units (SI) . Wherever possible, use the SI. Here are some other tips for formatting units of measurement:

  • Add spaces before units of measurement. For example, 2.5 mL not 2.5mL.
  • Be consistent with your units of measure (especially date and time). For example, 3 hours or 3 h.

When presenting statistical information, you must provide enough specific information to accurately describe the relationships among your data. Nothing is more frustrating to a reviewer than vague sentences about a variable being significant without any supporting details. The author guidelines for the journal Nature recommend that the following be included for statistical testing: the name of each statistical analysis, along with its n value; an explanation of why the test was used and what is being compared; and the specific alpha levels and P values for each test.

Angel Borja, writing for Elsevier publications, described the statistical rules for article formatting as follows:

  • Indicate the statistical tests used with all relevant parameters.
  • Use mean and standard deviation to report normally distributed data.
  • Use median and interpercentile range to report skewed data.
  • For numbers, use two significant digits unless more precision is necessary.
  • Never use percentages for very small samples.

Remember, you must be prepared to justify your findings and conclusions, and one of the best ways to do this is through factual accuracy and the acknowledgment of opposing interpretations, data, and/or points of view.

Even though you may not look forward to the process of formatting your research paper, it's important to present your findings clearly, consistently, and professionally. With the right paper format, your chances of publication increase, and your research will be more likely to make an impact in your field. Don't underestimate the details. They are the backbone of scientific writing and research.

One last tip: Before you submit your research, consider using our academic editing service for expert help with paper formatting, editing, and proofreading. We can tailor your paper to specific journal guidelines at your request.

Image source: 85Fifteen/ Unsplash.com

Let Us Format Your Paper to Your Target Journal’s Guidelines

Hire an expert academic editor , or get a free sample, about the author.

Joanna's passion for English literature (proven by her M.A. thesis on Jane Austen) is matched by her passion to help others with their writing (shown by her role as an in-house editor with Scribendi). She enjoys lively discussions about plot, character, and nerdy TV shows with her husband, and she loves singing almost as much as she loves reading. Isn't music another language after all?

Have You Read?

"The Complete Beginner's Guide to Academic Writing"

Related Posts

APA Style and APA Formatting

APA Style and APA Formatting

How to Research a Term Paper

How to Research a Term Paper

How to Write a Thesis or Dissertation

How to Write a Thesis or Dissertation

Upload your file(s) so we can calculate your word count, or enter your word count manually.

We will also recommend a service based on the file(s) you upload.

English is not my first language. I need English editing and proofreading so that I sound like a native speaker.

I need to have my journal article, dissertation, or term paper edited and proofread, or I need help with an admissions essay or proposal.

I have a novel, manuscript, play, or ebook. I need editing, copy editing, proofreading, a critique of my work, or a query package.

I need editing and proofreading for my white papers, reports, manuals, press releases, marketing materials, and other business documents.

I need to have my essay, project, assignment, or term paper edited and proofread.

I want to sound professional and to get hired. I have a resume, letter, email, or personal document that I need to have edited and proofread.

 Prices include your personal % discount.

 Prices include % sales tax ( ).

scientific paper example

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

scientific paper example

Writing the Scientific Paper

When you write about scientific topics to specialists in a particular scientific field, we call that scientific writing. (When you write to non-specialists about scientific topics, we call that science writing.)

The scientific paper has developed over the past three centuries into a tool to communicate the results of scientific inquiry. The main audience for scientific papers is extremely specialized. The purpose of these papers is twofold: to present information so that it is easy to retrieve, and to present enough information that the reader can duplicate the scientific study. A standard format with six main part helps readers to find expected information and analysis:

  • Title--subject and what aspect of the subject was studied.
  • Abstract--summary of paper: The main reason for the study, the primary results, the main conclusions
  • Introduction-- why the study was undertaken
  • Methods and Materials-- how the study was undertaken
  • Results-- what was found
  • Discussion-- why these results could be significant (what the reasons might be for the patterns found or not found)

There are many ways to approach the writing of a scientific paper, and no one way is right. Many people, however, find that drafting chunks in this order works best: Results, Discussion, Introduction, Materials & Methods, Abstract, and, finally, Title.

The title should be very limited and specific. Really, it should be a pithy summary of the article's main focus.

  • "Renal disease susceptibility and hypertension are under independent genetic control in the fawn hooded rat"
  • "Territory size in Lincoln's Sparrows ( Melospiza lincolnii )"
  • "Replacement of deciduous first premolars and dental eruption in archaeocete whales"
  • "The Radio-Frequency Single-Electron Transistor (RF-SET): A Fast and Ultrasensitive Electrometer"

This is a summary of your article. Generally between 50-100 words, it should state the goals, results, and the main conclusions of your study. You should list the parameters of your study (when and where was it conducted, if applicable; your sample size; the specific species, proteins, genes, etc., studied). Think of the process of writing the abstract as taking one or two sentences from each of your sections (an introductory sentence, a sentence stating the specific question addressed, a sentence listing your main techniques or procedures, two or three sentences describing your results, and one sentence describing your main conclusion).

Example One

Hypertension, diabetes and hyperlipidemia are risk factors for life-threatening complications such as end-stage renal disease, coronary artery disease and stroke. Why some patients develop complications is unclear, but only susceptibility genes may be involved. To test this notion, we studied crosses involving the fawn-hooded rat, an animal model of hypertension that develops chronic renal failure. Here, we report the localization of two genes, Rf-1 and Rf-2 , responsible for about half of the genetic variation in key indices of renal impairment. In addition, we localize a gene, Bpfh-1 , responsible for about 26% of the genetic variation in blood pressure. Rf-1 strongly affects the risk of renal impairment, but has no significant effect on blood pressure. Our results show that susceptibility to a complication of hypertension is under at least partially independent genetic control from susceptibility to hypertension itself.

Brown, Donna M, A.P. Provoost, M.J. Daly, E.S. Lander, & H.J. Jacob. 1996. "Renal disease susceptibility and hypertension are under indpendent genetic control in the faun-hooded rat." Nature Genetics , 12(1):44-51.

Example Two

We studied survival of 220 calves of radiocollared moose ( Alces alces ) from parturition to the end of July in southcentral Alaska from 1994 to 1997. Prior studies established that predation by brown bears ( Ursus arctos ) was the primary cause of mortality of moose calves in the region. Our objectives were to characterize vulnerability of moose calves to predation as influenced by age, date, snow depths, and previous reproductive success of the mother. We also tested the hypothesis that survival of twin moose calves was independent and identical to that of single calves. Survival of moose calves from parturition through July was 0.27 ± 0.03 SE, and their daily rate of mortality declined at a near constant rate with age in that period. Mean annual survival was 0.22 ± 0.03 SE. Previous winter's snow depths or survival of the mother's previous calf was not related to neonatal survival. Selection for early parturition was evidenced in the 4 years of study by a 6.3% increase in the hazard of death with each daily increase in parturition date. Although there was no significant difference in survival of twin and single moose calves, most twins that died disappeared together during the first 15 days after birth and independently thereafter, suggesting that predators usually killed both when encountered up to that age.

Key words: Alaska, Alces alces , calf survival, moose, Nelchina, parturition synchrony, predation

Testa, J.W., E.F. Becker, & G.R. Lee. 2000. "Temporal patterns in the survival of twin and single moose ( alces alces ) calves in southcentral Alaska." Journal of Mammalogy , 81(1):162-168.

Example Three

We monitored breeding phenology and population levels of Rana yavapaiensis by use of repeated egg mass censuses and visual encounter surveys at Agua Caliente Canyon near Tucson, Arizona, from 1994 to 1996. Adult counts fluctuated erratically within each year of the study but annual means remained similar. Juvenile counts peaked during the fall recruitment season and fell to near zero by early spring. Rana yavapaiensis deposited eggs in two distinct annual episodes, one in spring (March-May) and a much smaller one in fall (September-October). Larvae from the spring deposition period completed metamorphosis in earlv summer. Over the two years of study, 96.6% of egg masses successfully produced larvae. Egg masses were deposited during periods of predictable, moderate stream flow, but not during seasonal periods when flash flooding or drought were likely to affect eggs or larvae. Breeding phenology of Rana yavapaiensis is particularly well suited for life in desert streams with natural flow regimes which include frequent flash flooding and drought at predictable times. The exotic predators of R. yavapaiensis are less able to cope with fluctuating conditions. Unaltered stream flow regimes that allow natural fluctuations in stream discharge may provide refugia for this declining ranid frog from exotic predators by excluding those exotic species that are unable to cope with brief flash flooding and habitat drying.

Sartorius, Shawn S., and Philip C. Rosen. 2000. "Breeding phenology of the lowland leopard frog ( Rana yavepaiensis )." Southwestern Naturalist , 45(3): 267-273.

Introduction

The introduction is where you sketch out the background of your study, including why you have investigated the question that you have and how it relates to earlier research that has been done in the field. It may help to think of an introduction as a telescoping focus, where you begin with the broader context and gradually narrow to the specific problem addressed by the report. A typical (and very useful) construction of an introduction proceeds as follows:

"Echimyid rodents of the genus Proechimys (spiny rats) often are the most abundant and widespread lowland forest rodents throughout much of their range in the Neotropics (Eisenberg 1989). Recent studies suggested that these rodents play an important role in forest dynamics through their activities as seed predators and dispersers of seeds (Adler and Kestrell 1998; Asquith et al 1997; Forget 1991; Hoch and Adler 1997)." (Lambert and Adler, p. 70)

"Our laboratory has been involved in the analysis of the HLA class II genes and their association with autoimmune disorders such as insulin-dependent diabetes mellitus. As part of this work, the laboratory handles a large number of blood samples. In an effort to minimize the expense and urgency of transportation of frozen or liquid blood samples, we have designed a protocol that will preserve the integrity of lymphocyte DNA and enable the transport and storage of samples at ambient temperatures." (Torrance, MacLeod & Hache, p. 64)

"Despite the ubiquity and abundance of P. semispinosus , only two previous studies have assessed habitat use, with both showing a generalized habitat use. [brief summary of these studies]." (Lambert and Adler, p. 70)

"Although very good results have been obtained using polymerase chain reaction (PCR) amplification of DNA extracted from dried blood spots on filter paper (1,4,5,8,9), this preservation method yields limited amounts of DNA and is susceptible to contamination." (Torrance, MacLeod & Hache, p. 64)

"No attempt has been made to quantitatively describe microhabitat characteristics with which this species may be associated. Thus, specific structural features of secondary forests that may promote abundance of spiny rats remains unknown. Such information is essential to understand the role of spiny rats in Neotropical forests, particularly with regard to forest regeneration via interactions with seeds." (Lambert and Adler, p. 71)

"As an alternative, we have been investigating the use of lyophilization ("freeze-drying") of whole blood as a method to preserve sufficient amounts of genomic DNA to perform PCR and Southern Blot analysis." (Torrance, MacLeod & Hache, p. 64)

"We present an analysis of microhabitat use by P. semispinosus in tropical moist forests in central Panama." (Lambert and Adler, p. 71)

"In this report, we summarize our analysis of genomic DNA extracted from lyophilized whole blood." (Torrance, MacLeod & Hache, p. 64)

Methods and Materials

In this section you describe how you performed your study. You need to provide enough information here for the reader to duplicate your experiment. However, be reasonable about who the reader is. Assume that he or she is someone familiar with the basic practices of your field.

It's helpful to both writer and reader to organize this section chronologically: that is, describe each procedure in the order it was performed. For example, DNA-extraction, purification, amplification, assay, detection. Or, study area, study population, sampling technique, variables studied, analysis method.

Include in this section:

  • study design: procedures should be listed and described, or the reader should be referred to papers that have already described the used procedure
  • particular techniques used and why, if relevant
  • modifications of any techniques; be sure to describe the modification
  • specialized equipment, including brand-names
  • temporal, spatial, and historical description of study area and studied population
  • assumptions underlying the study
  • statistical methods, including software programs

Example description of activity

Chromosomal DNA was denatured for the first cycle by incubating the slides in 70% deionized formamide; 2x standard saline citrate (SSC) at 70ºC for 2 min, followed by 70% ethanol at -20ºC and then 90% and 100% ethanol at room temperature, followed by air drying. (Rouwendal et al ., p. 79)

Example description of assumptions

We considered seeds left in the petri dish to be unharvested and those scattered singly on the surface of a tile to be scattered and also unharvested. We considered seeds in cheek pouches to be harvested but not cached, those stored in the nestbox to be larderhoarded, and those buried in caching sites within the arena to be scatterhoarded. (Krupa and Geluso, p. 99)

Examples of use of specialized equipment

  • Oligonucleotide primers were prepared using the Applied Biosystems Model 318A (Foster City, CA) DNA Synthesizer according to the manufacturers' instructions. (Rouwendal et al ., p.78)
  • We first visually reviewed the complete song sample of an individual using spectrograms produced on a Princeton Applied Research Real Time Spectrum Analyzer (model 4512). (Peters et al ., p. 937)

Example of use of a certain technique

Frogs were monitored using visual encounter transects (Crump and Scott, 1994). (Sartorius and Rosen, p. 269)

Example description of statistical analysis

We used Wilcox rank-sum tests for all comparisons of pre-experimental scores and for all comparisons of hue, saturation, and brightness scores between various groups of birds ... All P -values are two-tailed unless otherwise noted. (Brawner et al ., p. 955)

This section presents the facts--what was found in the course of this investigation. Detailed data--measurements, counts, percentages, patterns--usually appear in tables, figures, and graphs, and the text of the section draws attention to the key data and relationships among data. Three rules of thumb will help you with this section:

  • present results clearly and logically
  • avoid excess verbiage
  • consider providing a one-sentence summary at the beginning of each paragraph if you think it will help your reader understand your data

Remember to use table and figures effectively. But don't expect these to stand alone.

Some examples of well-organized and easy-to-follow results:

  • Size of the aquatic habitat at Agua Caliente Canyon varied dramatically throughout the year. The site contained three rockbound tinajas (bedrock pools) that did not dry during this study. During periods of high stream discharge seven more seasonal pools and intermittent stretches of riffle became available. Perennial and seasonal pool levels remained stable from late February through early May. Between mid-May and mid-July seasonal pools dried until they disappeared. Perennial pools shrank in surface area from a range of 30-60 m² to 3-5- M². (Sartorius and Rosen, Sept. 2000: 269)

Notice how the second sample points out what is important in the accompanying figure. It makes us aware of relationships that we may not have noticed quickly otherwise and that will be important to the discussion.

A similar test result is obtained with a primer derived from the human ß-satellite... This primer (AGTGCAGAGATATGTCACAATG-CCCC: Oligo 435) labels 6 sites in the PRINS reaction: the chromosomes 1, one pair of acrocentrics and, more weakly, the chromosomes 9 (Fig. 2a). After 10 cycles of PCR-IS, the number of sites labeled has doubled (Fig. 2b); after 20 cycles, the number of sites labeled is the same but the signals are stronger (Fig. 2c) (Rouwendal et al ., July 93:80).

Related Information: Use Tables and Figures Effectively

Do not repeat all of the information in the text that appears in a table, but do summarize it. For example, if you present a table of temperature measurements taken at various times, describe the general pattern of temperature change and refer to the table.

"The temperature of the solution increased rapidly at first, going from 50º to 80º in the first three minutes (Table 1)."

You don't want to list every single measurement in the text ("After one minute, the temperature had risen to 55º. After two minutes, it had risen to 58º," etc.). There is no hard and fast rule about when to report all measurements in the text and when to put the measurements in a table and refer to them, but use your common sense. Remember that readers have all that data in the accompanying tables and figures, so your task in this section is to highlight key data, changes, or relationships.

In this section you discuss your results. What aspect you choose to focus on depends on your results and on the main questions addressed by them. For example, if you were testing a new technique, you will want to discuss how useful this technique is: how well did it work, what are the benefits and drawbacks, etc. If you are presenting data that appear to refute or support earlier research, you will want to analyze both your own data and the earlier data--what conditions are different? how much difference is due to a change in the study design, and how much to a new property in the study subject? You may discuss the implication of your research--particularly if it has a direct bearing on a practical issue, such as conservation or public health.

This section centers on speculation . However, this does not free you to present wild and haphazard guesses. Focus your discussion around a particular question or hypothesis. Use subheadings to organize your thoughts, if necessary.

This section depends on a logical organization so readers can see the connection between your study question and your results. One typical approach is to make a list of all the ideas that you will discuss and to work out the logical relationships between them--what idea is most important? or, what point is most clearly made by your data? what ideas are subordinate to the main idea? what are the connections between ideas?

Achieving the Scientific Voice

Eight tips will help you match your style for most scientific publications.

  • Develop a precise vocabulary: read the literature to become fluent, or at least familiar with, the sort of language that is standard to describe what you're trying to describe.
  • Once you've labeled an activity, a condition, or a period of time, use that label consistently throughout the paper. Consistency is more important than creativity.
  • Define your terms and your assumptions.
  • Include all the information the reader needs to interpret your data.
  • Remember, the key to all scientific discourse is that it be reproducible . Have you presented enough information clearly enough that the reader could reproduce your experiment, your research, or your investigation?
  • When describing an activity, break it down into elements that can be described and labeled, and then present them in the order they occurred.
  • When you use numbers, use them effectively. Don't present them so that they cause more work for the reader.
  • Include details before conclusions, but only include those details you have been able to observe by the methods you have described. Do not include your feelings, attitudes, impressions, or opinions.
  • Research your format and citations: do these match what have been used in current relevant journals?
  • Run a spellcheck and proofread carefully. Read your paper out loud, and/ or have a friend look over it for misspelled words, missing words, etc.

Applying the Principles, Example 1

The following example needs more precise information. Look at the original and revised paragraphs to see how revising with these guidelines in mind can make the text clearer and more informative:

Before: Each male sang a definite number of songs while singing. They start with a whistle and then go from there. Each new song is always different, but made up an overall repertoire that was completed before starting over again. In 16 cases (84%), no new songs were sung after the first 20, even though we counted about 44 songs for each bird.
After: Each male used a discrete number of song types in his singing. Each song began with an introductory whistle, followed by a distinctive, complex series of fluty warbles (Fig. 1). Successive songs were always different, and five of the 19 males presented their entire song repertoire before repeating any of their song types (i.e., the first IO recorded songs revealed the entire repertoire of 10 song types). Each song type recurred in long sequences of singing, so that we could be confident that we had recorded the entire repertoire of commonly used songs by each male. For 16 of the 19 males, no new song types were encountered after the first 20 songs, even though we analyzed and average of 44 songs/male (range 30-59).

Applying the Principles, Example 2

In this set of examples, even a few changes in wording result in a more precise second version. Look at the original and revised paragraphs to see how revising with these guidelines in mind can make the text clearer and more informative:

Before: The study area was on Mt. Cain and Maquilla Peak in British Columbia, Canada. The study area is about 12,000 ha of coastal montane forest. The area is both managed and unmanaged and ranges from 600-1650m. The most common trees present are mountain hemlock ( Tsuga mertensiana ), western hemlock ( Tsuga heterophylla ), yellow cedar ( Chamaecyparis nootkatensis ), and amabilis fir ( Abies amabilis ).
After: The study took place on Mt. Cain and Maquilla Peak (50'1 3'N, 126'1 8'W), Vancouver Island, British Columbia. The study area encompassed 11,800 ha of coastal montane forest. The landscape consisted of managed and unmanaged stands of coastal montane forest, 600-1650 m in elevation. The dominant tree species included mountain hemlock ( Tsuga mertensiana ), western hemlock ( Tsuga heterophylla ), yellow cedar ( Chamaecyparis nootkatensis ), and amabilis fir ( Abies amabilis ).

Two Tips for Sentence Clarity

Although you will want to consider more detailed stylistic revisions as you become more comfortable with scientific writing, two tips can get you started:

First, the verb should follow the subject as soon as possible.

Really Hard to Read : "The smallest of the URF's (URFA6L), a 207-nucleotide (nt) reading frame overlapping out of phase the NH2- terminal portion of the adenosinetriphosphatase (ATPase) subunit 6 gene has been identified as the animal equivalent of the recently discovered yeast H+-ATPase subunit gene."

Less Hard to Read : "The smallest of the UR-F's is URFA6L, a 207-nucleotide (nt) reading frame overlapping out of phase the NH2-terminal portion of the adenosinetriphosphatase (ATPase) subunit 6 gene; it has been identified as the animal equivalent of the recently discovered yeast H+-ATPase subunit 8 gene."

Second, place familiar information first in a clause, a sentence, or a paragraph, and put the new and unfamiliar information later.

More confusing : The epidermis, the dermis, and the subcutaneous layer are the three layers of the skin. A layer of dead skin cells makes up the epidermis, which forms the body's shield against the world. Blood vessels, carrying nourishment, and nerve endings, which relay information about the outside world, are found in the dermis. Sweat glands and fat cells make up the third layer, the subcutaneous layer.

Less confusing : The skin consists of three layers: the epidermis, the dermis, and the subcutaneous layer. The epidermis is made up of dead skin cells, and forms a protective shield between the body and the world. The dermis contains the blood vessels and nerve endings that nourish the skin and make it receptive to outside stimuli. The subcutaneous layer contains the sweat glands and fat cells which perform other functions of the skin.

Bibliography

  • Scientific Writing for Graduate Students . F. P. Woodford. Bethesda, MD: Council of Biology Editors, 1968. [A manual on the teaching of writing to graduate students--very clear and direct.]
  • Scientific Style and Format . Council of Biology Editors. Cambridge: Cambridge University Press, 1994.
  • "The science of scientific writing." George Gopen and Judith Swann. The American Scientist , Vol. 78, Nov.-Dec. 1990. Pp 550-558.
  • "What's right about scientific writing." Alan Gross and Joseph Harmon. The Scientist , Dec. 6 1999. Pp. 20-21.
  • "A Quick Fix for Figure Legends and Table Headings." Donald Kroodsma. The Auk , 117 (4): 1081-1083, 2000.

Wortman-Wunder, Emily, & Kate Kiefer. (1998). Writing the Scientific Paper. Writing@CSU . Colorado State University. https://writing.colostate.edu/resources/writing/guides/.

UCI Libraries Mobile Site

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

Libaries home page

Email this link

Writing a scientific paper.

  • Writing a lab report

What is a "good" introduction?

Citing sources in the introduction, "introduction checklist" from: how to write a good scientific paper. chris a. mack. spie. 2018..

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

This is where you describe briefly and clearly why you are writing the paper. The introduction supplies sufficient background information for the reader to understand and evaluate the experiment you did. It also supplies a rationale for the study.

  • Present the problem and the proposed solution
  • Presents nature and scope of the problem investigated
  • Reviews the pertinent literature to orient the reader
  • States the method of the experiment
  • State the principle results of the experiment

It is important to cite sources in the introduction section of your paper as evidence of the claims you are making. There are ways of citing sources in the text so that the reader can find the full reference in the literature cited section at the end of the paper, yet the flow of the reading is not badly interrupted. Below are some example of how this can be done:     "Smith (1983) found that N-fixing plants could be infected by several different species of Rhizobium."     "Walnut trees are known to be allelopathic (Smith 1949,  Bond et al. 1955, Jones and Green 1963)."     "Although the presence of Rhizobium normally increases the growth of legumes (Nguyen 1987), the opposite effect has been observed (Washington 1999)." Note that articles by one or two authors are always cited in the text using their last names. However, if there are more than two authors, the last name of the 1st author is given followed by the abbreviation et al. which is Latin for "and others". 

From:  https://writingcenter.gmu.edu/guides/imrad-reports-introductions

  • Indicate the field of the work, why this field is important, and what has already been done (with proper citations).
  • Indicate a gap, raise a research question, or challenge prior work in this territory.
  • Outline the purpose and announce the present research, clearly indicating what is novel and why it is significant.
  • Avoid: repeating the abstract; providing unnecessary background information; exaggerating the importance of the work; claiming novelty without a proper literature search. 
  • << Previous: ABSTRACT
  • Next: METHODS >>
  • 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.

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.

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.
  • 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

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
  • Blood Transfus
  • v.11(2); 2013 Apr

How to write a scientific manuscript for publication

Giancarlo maria liumbruno.

1 Immunohaematology, Transfusion Medicine and Clinical Pathology Units, “San Giovanni Calibita” Fatebenefratelli Hospital, AFAR, Rome, Italy

Claudio Velati

2 Immunohaematology and Transfusion Medicine Department, Ospedale Maggiore Pizzardi, Azienda USL Bologna, Bologna, Italy

Patrizio Pasqualetti

3 Medical Statistics & Information Technology, Fatebenefratelli Association for Research, Isola Tiberina, Rome, Italy

Massimo Franchini

4 Department of Transfusion Medicine and Haematology, Carlo Poma Hospital, Mantua, Italy

Introduction

The origins and development of the scientific and technical press can be traced back to 1665 when the first “modern” scientific papers appeared and were characterized by non standardised form and style 1 . Subsequently, nearly 300 years ago 2 , in an attempt to ensure that articles met the journal’s standards of quality and scientific validity, the peer-reviewed process for scientific manuscripts was born in England and France. Since then, there has been an enormous proliferation of scientific journals and manuscripts so that, at present, the numbers of biomedical papers published annually by over 20,000 journals, at a rate of 5,500 new papers per day, far exceeds 2,000,000 1 , 2 .

Published scientific papers and professional meetings are really essential to disseminate relevant information and research findings. However, most of the abstracts of presentations given at scientific meetings are usually available only in conference proceedings although they have the potential to be subsequently published as articles in peer-reviewed journals.

A recently published Cochrane review showed that only 44.5% of almost 30,000 scientific meeting abstracts were published as articles 3 . No association between full publication and authors’ country of origin was detected. Factors associated with full publication included acceptance vs rejection of abstracts for oral or poster presentations, acceptance for oral presentations rather than poster sessions, “positive” results, using the report authors’ definition of “positive”, randomised trial study design and basic rather than clinical research.

Possible reasons for failed publication include lack of time, research still underway, problems with co-authors and negative results 4 . Undoubtedly, lack of the necessary skills and experience in the process of writing and publishing is another possible contributing factor also in the field of Transfusion Medicine although the specialists in this discipline are currently adopting the principles and research methodologies that support evidence-based medicine 5 , and high-level research is actually being carried out at the same rate as in all medical specialties.

There are three broad groups of manuscripts: original scientific articles, reviews and case reports. Although case reports are part of the evidence hierarchy in evidence-based practice, albeit at a lower level, and case series are incorporated in a significant proportion of health technology assessments 6 , this article will address the multiple steps required in writing original articles and reviews with the aim of providing the reader with the necessary tools to prepare, submit and successfully publish a manuscript.

The anatomy of a paper: from origin to current format

The history of scientific journals dates from 1665, when the French “Journal des sçavans” and the English “Philosophical Transactions of the Royal Society” first began systematically publishing research results 7 . From then on, the initial structure of scientific papers evolved gradually from letters (usually by a single author, with a polite style and contemporarily addressing multiple subjects) and experimental reports (essentially descriptive and presenting experiences and effects in chronological order) to a better structured and more fluent form characterised by an embryonic description of methods and interpretation of results. This evolved way of reporting experiments gradually replaced the letter form.

It was not, however, until the second half of the 19 th century that the method description became fully developed and a comprehensive organisation of the manuscripts known as “theory-experiment-discussion” emerged 1 . At the beginning of the last century a gradual decrease of the use of the literary style coincided with a growing standardisation of the editorial rules that paved the way for the formal established Introduction, Methods, Results, and Discussion (IMRAD) structure of scientific papers, which was adopted in the 1980s.

At present, IMRAD is the format encouraged for the text of observational (i.e. retrospective/descriptive) and experimental (i.e. randomised controlled) studies by the “Uniform Requirements for Manuscripts Submitted to Biomedical Journals” which have become the most important and widely accepted (by over 500 biomedical journals) guide to writing, publishing, and editing in international biomedical publications 8 . The Uniform Requirements are released by the International Committee of Medical Journal Editors (ICMJE), an evolution of the initial group of Journal Editors who met for the first time in Vancouver in 1978 and subsequently issued a number of editorial policy statements and guidelines for manuscript submission.

According to the ICMJE, “this so-called IMRAD structure is not an arbitrary publication format but rather a direct reflection of the process of scientific discovery” 9 . In addition it facilitates modular reading and locating of specific information, which is normally found in pre-established sections of an article 7 .

“Long articles may need subheadings within some sections (especially Results and Discussion) to clarify their content. Other types of articles, such as case reports, reviews and editorials, probably need to be formatted differently” 9 .

This format does not comprise other important and integral parts of the article, such as the Title Page, Abstract, Acknowledgements, Figures and Tables (comprising their legends) and References 8 .

There are often slight variations from one journal’s format to another but every journal has instructions to authors available on their website and it is crucial that authors download and comply with them.

The latest edition of the Uniform Requirements was updated in April 2010; it is available at the ICMJE website and is an essential guideline for all authors writing a biomedical manuscript 9 .

Consolidated standards of reporting trials

Medical science depends entirely on the transparent reporting of clinical trials 10 .

Unfortunately, several reviews have documented deficiencies in reports of clinical trials 11 – 15 .

In 1996, a group of scientists and editors developed the CONsolidated Standards Of Reporting Trials (CONSORT) statement which is intended to improve the reporting of a randomised, controlled trial (RCT), enabling readers to understand the design of a trial, its conduct, analysis and interpretation and to assess the validity of its results 16 . It emphasises that this can only be achieved through complete transparency from authors.

The CONSORT statement was updated in 2001 and after the 2007 meeting the statement was further revised and published as CONSORT 2010 which is the most up-to-date version and can be freely viewed and downloaded through one of the several link to Journals available at the CONSORT website under the section “CONSORT Statement - Downloads” 17 . The statement facilitates critical appraisal and interpretation of RCT and many leading medical journals and major international editorial groups have endorsed it.

The statement consists of a checklist (25 items) and a flow diagram that authors can use for reporting a RCT. The checklist items pertain to the content of the Title, Abstract, Introduction, Methods, Results, Discussion and Other information. The flow diagram is intended to depict the passage of participants through a RCT (enrolment, intervention allocation, follow-up and analysis). It is strongly recommended that the CONSORT Statement be used in conjunction with the CONSORT Explanation and Elaboration Document which is available at the CONSORT website under the above mentioned section 17 .

Another major point to consider is the obligation to register clinical trials 9 .

In September 2004 the ICMJE changed their policy and decided they would consider trials for publication only if they had been registered before the enrolment of the first participant. The ICMJE accepts registration in the international registries listed in Table I .

International trial registries acceptable to the International Committee of Medical Journal Editors and relevant websites.

Strengthening the reporting of observational studies in epidemiology

The reporting of observational studies frequently lacks details and is not clear enough 18 , 19 . Consequently the quality is poor although many questions in medical research are investigated in observational studies and overwhelming evidence is also extrapolated from them 20 . In fact, observational studies are more suitable for the detection of rare or late adverse effects of treatments, and are more likely to provide an indication of what is achieved in daily medical practice 21 .

To improve the reporting of observational studies (cohort, case-control or cross-sectional studies) a group of methodologists, researchers and editors developed a useful checklist of 22 items: the StrengThening the Reporting of OBservational studies in Epidemiology (STROBE) Statement 21 . The checklist items pertain to the content of the Title, Abstract, Introduction, Methods, Results, Discussion and Other information sections of articles. The STROBE checklists can be freely viewed and downloaded at the STROBE website under the section “Available checklists” 22 . They also include a draft checklist for conference abstracts (items to be included when reporting observational studies in a conference abstract) pertaining to the content of the following sections: Title, Authors, Study design, Objective, Methods, Results and Conclusion.

The STROBE Statement provides guidance to authors on how to improve the reporting of observational studies, it facilitates critical appraisal and interpretation of studies and is widely supported by reviewers, a growing number of biomedical journal editors and readers.

The STROBE checklist is best used in conjunction with an explanation and elaboration article which discusses each of the 22 checklist items, gives methodological background, publishes examples of transparent reporting and is freely available at the STROBE Statement website under the above mentioned section through the link with the Journals in which the document has been published (PLoS Medicine, Annals of Internal Medicine and Epidemiology) 22 .

As review articles comprehensively cover a specific biomedical topic and justify future research directions, they require that the author extensively review and master the literature and then develop some general statements and conclusions with practical implications for patients’ care 23 , 24 . In addition, they should provide an updated reference for those readers interested in broadening their knowledge of critical issues. Review articles are, therefore, important not only for younger physicians early in their career but also for senior academic staff as they represent a tool for intellectual enrichment and enhancement of the standards of research. Writing a review requires knowledge and continuous improvement of qualifications in line with the accumulation of better and updated scientific literature evidence. For this reason, journals often invite experts on a specific topic to write a review article. However, authors can also ask Editors if they would be interested in publishing a review article on a particular, topical, relevant and debated issue.

As reviews are the most accessed among the various types of articles and contribute substantially to the impact factor of journals, obviously they are welcomed and encouraged by many journals and have become an inseparable part of the writing scientific culture.

The three basic types of literature reviews are narrative reviews (which include editorials, commentaries and narrative overviews or non-systematic narrative reviews), qualitative systematic reviews and quantitative systematic reviews (meta-analyses) ( Table II ) 25 .

Summary of the types of literature reviews.

Editorials, typically written by the editor of the journal or an invited guest, may be a narrative review if the author retrieves and summarises information about a particular topic for the reader 25 . Usually, these types of narrative reviews are based upon a short, select and narrowly focused review of only a few papers. However, editorials may be no more than the editor’s comments regarding a current issue of the journal or a current event in health care and do not, therefore, automatically qualify as narrative reviews.

Commentaries

Commentaries may also be written as a narrative review; however, they are typically written with a particular opinion being expressed 25 . Research methodology is not usually presented in these articles which reflect the author’s biased synthesis of other articles. Commentaries are usually shorter than a full-length review article and the author should be an expert in the content area of the commentary. Usually, the purpose of a commentary is to stimulate academic debate between the journal’s readers.

Narrative reviews

Non-systematic narrative reviews are comprehensive narrative syntheses of previously published information 26 . This type of literature review reports the author’s findings in a condensed format that typically summarises the contents of each article. Authors of narrative overviews are often acknowledged experts in the field and have conducted research themselves. Editors sometimes solicit narrative overviews from specific authors in order to bring certain issues to light. Although the bibliographic research methodology is an obligatory section in systematic reviews and meta-analyses, it is also becoming an inseparable part of narrative literature reviews. Providing information on the databases accessed, terms, inclusion and exclusion criteria and time limits adds objectivity to the main messages and conclusions. It is advisable to use only credible databases (at least two or three) which only select high-quality publications that contain the most up-to-date information (see Table III ) 24 . The best way to organise the analysis of the sources in the main text of a narrative biomedical review is to transform information from the retrieved publications into bibliographic cards with a short description of the main results, level of evidence, strengths and limitations of each study and relevance to each section of the manuscript. Furthermore, the readability of a review can be improved by including a few self-explanatory tables, boxes, and figures synthesising essential information and conveying original messages 24 . We also suggest the use of software packages for reference management, which saves time during the multiple revisions.

Main online libraries, catalogues and databases.

In conclusion, a successful narrative review should have the following characteristics: be well-structured, synthesise the available evidence pertaining to the topic, convey a clear message and draw conclusions supported by data analysis.

Qualitative systematic reviews

Qualitative systematic reviews are a type of literature review that employ detailed, rigorous and explicit methods and are, therefore, a more powerful evidence-based source to garner clinical information than narrative reviews, case reports, case series, and poorly conducted cohort studies. A detailed bibliographic research based upon a focused question or purpose is the peculiar characteristic of a systematic review 27 . These reviews are called qualitative because the process by which the individual studies are integrated includes a summary and critique of the findings derived from systematic methods, but does not statistically combine the results of all of the studies reviewed.

Quantitative systematic reviews

A quantitative systematic review or meta-analysis critically evaluates each paper and statistically combines the results of the studies 28 . The authors of a meta-analysis employ all of the rigorous methodology of qualitative systematic reviews and, in addition, gather the original patients’ data from each of the studies under review, pool it all together in a database and produce the appropriate statistics on this larger sample. While this process leads to a more powerful and generalizable conclusion, which is the strength of the meta-analysis, on the other hand it can pool together studies that are very heterogeneous which is the main drawback of a quantitative systematic review. Nevertheless, well-executed quantitative systematic reviews constitute the highest level of evidence for medical decision making 28 .

The recently published Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA) statement aims to help improve reporting, focusing on systematic reviews of RCT. The Statement consists of a checklist of 27 essential items for transparent reporting and a flow diagram for the phases of study selection and is accompanied by the PRISMA Explanation and Elaboration Document, which, among other things, provides examples of good reporting for the various review sections 29 .

A further guidance on the reporting of systematic reviews has been published by the Cochrane Collaboration, an international organisation that prepares, updates and publishes systematic reviews of the effects of health-care interventions following a standardised format 30 .

Preparing to write a manuscript Background information

The question or hypothesis formulated by the investigator is the common starting point to search the relevant published literature for an answer 31 . Gathering the background information through an extensive literature search relevant to the topic of interest is the subsequent essential step. Peer reviewers are often experts and not citing important articles poses the manuscript at risk of rejection. It is advisable to consult at least two or three credible databases (see Table III ) to identify the crucial relevant articles and to track down “landmark” articles. In addition, avoid using papers published more than 10 years ago and do not rely on just the abstracts but obtain full-text articles. Articles relevant to the research topic and published in the journal in which the paper is to be submitted should be reviewed and cited 32 .

Last but not least, the bibliographical search should also aim at finding recently published articles similar to the one the author intends to submit. In fact, a journal can be less interested in publishing such a manuscript unless the results reflect new or different findings.

Target journal

It can be worth thinking about this issue before starting to write as a proper choice of the journal can affect not only the writing style but also the ease of publication and the prompt dissemination of research. Ideally, the target journal should be the one in which similar work has been published 32 .

Electronic and open-access journals are the latest resources for publishing and data dissemination available on the scientific journal horizon.

It is also worth considering an appropriate level of impact factor or journal quality. The impact factor of a journal is a measure reflecting the average number of citations to recent articles published in science and social science journals. It is determined by the ratio of the number of citations of papers from that journal in the whole of the biomedical literature over a 2-year period. It is frequently used as a proxy for the relative importance of a journal within its field, with journals with higher impact factors deemed to be more important than those with lower ones.

It is also extremely important to read the instructions to authors section of the selected journal carefully. In fact, although there is a general style for most biomedical journals as agreed by the ICMJE in the Uniform Requirements 9 , individual journals may differ slightly in detail.

It is always best to sort out authorship before writing a manuscript as authorship order can be a source of problems once the paper has been written 23 .

Several guidelines relating to authorship are available and this issue has been extensively addressed in a recently published review article by Elizabeth Wager 33 . Most guidelines on the authorship of scientific articles are focused more on creative and intellectual aspects of research than on routine or technical contributions.

Alhough not universally accepted, the authorship criteria suggested by the ICMJE are the ones most widely promoted by medical journals 9 . According to these criteria, co-authors should: (i) substantially contribute to conception and design of the study, acquisition of data, or analysis and interpretation of data; (ii) draft the article or revise it critically for important intellectual content; and (iii) approve the final version.

The authors are listed in decreasing order of their contribution and the senior author, or mentor, should be the last but this convention has never been codified 33 .

It is advisable to provide accurate affiliations and contacts as they will be published on PubMed as well as in the journal but it is also important to agree on the corresponding author who should have full access to the study data and through the provided e-mail address will be the link with the scientific community for the future 1 .

Ethical issues

In addition to the authorship discussed above, there are several ethical issues involved in writing a paper. These include fabrication of data, duplicate publication, plagiarism, misuse of statistics, manipulation of images and inadequate or obviously false citations 31 .

A must-read for all those who are involved in any editorial activity are the guidelines released by the Committee on Publication Ethics (COPE) which is a forum for editors and publishers of peer-reviewed journals to discuss all aspects of publication ethics 34 . COPE provides advice to editors and publishers on all aspects of publication ethics and, in particular, how to handle cases of research and publication misconduct.

Writing the manuscript

Several models for the initial draft exist. A useful algorithm for writing a scientific manuscript is the one recently published by O’Connor and Holmquist 35 . According to these authors, the writing should start with making figures and tables, and then proceed with summary statements (the conclusions summarising the major contributions of the manuscript to the scientific community), identification of the audience, materials and methods, results, discussion, references, introduction, title and conclusion. The aim of this algorithm is to give the structural backbone to the manuscript and is designed to overcome writer’s block and to assist scientists who are not native English speakers.

A further and more general strategy to increase productivity during the early phases of manuscript writing is to ignore at the outset all the details that can be approached later such as structure, grammar and spelling.

The sequence of writing should address the following core sections of the paper in the order from first to last: methods, results, discussion and introduction 31 , 36 , 37 .

“Like every well-written story, a scientific manuscript should have a beginning (Introduction), middle (Materials and Methods), and an end (Results). The Discussion (the moral of the story) puts the study in perspective. The Abstract is an opening summary of the story and the Title gives the story a name” 38 . However, as correctly pointed out by Michael McKay, “writing is not necessarily in the temporal order of the final document (i.e. the IMRAD format)” 39 .

The take-home messages are, therefore: (i) a clear understanding of the essential components of each of these sections is critical to the successful composition of a scientific manuscript; (ii) the proper order of writing greatly facilitates the ease of writing; (iii) the approach to writing can be customised by authors on the basis both of the subject they are dealing with and their personal experience; (iv) the CONSORT 16 , 17 , STROBE 21 , 22 or PRISMA 29 statement must be used as a guidance document for the appropriate reporting of the type of study the authors are dealing with 31 , 32 , 38 .

In the following part of this paper the different sections of a manuscript will be dealt with in the order they are presented in the final document.

Title, keywords and abstract

The title is determinant for the indexing process of the article and greatly contributes to the visibility of the paper. It should reflect the essence of the article, its novelty and its relevance to the biomedical field it deals with 24 . It should be clear, brief, specific, not include jargon or non-standard and unexplained abbreviations, reflect the purpose of the study and state the issue(s) addressed rather than the conclusions 38 . Indicative titles are, therefore, better than declarative ones. Obviously, the title and abstract should correlate with each other.

Available evidence suggests that the presence of a colon in the title positively correlates with the number of citations 40 . In other words, the more specific and accurate the description of the content is, the more chance the manuscript has of being cited 38 .

The title of systematic reviews should ideally follow the participants, interventions, comparisons, outcomes, and study design (PICOS) approach, and include the terms “systematic review”, “meta-analysis”, or both 41 .

The keywords enable the database searching of the article and should be provided in compliance with the instructions to authors. A careful choice from the Medical Subject Headings (MeSH) in the National Library of Medicine (NLM) controlled vocabulary thesaurus used for indexing articles in PubMed greatly increases the chances the paper is retrieved and cited by other authors 42 .

The abstract is the last section to be written but it is the most important part of a paper because it is usually the first to be read and readers use the information contained in it to decide whether to read the whole article or not. It should be a concise summary of the manuscript and no longer than specified in the instructions to authors. Usually, abstracts do not contain references and abbreviations and acronyms are not always allowed. If required, it has to be structured in a specific way. For example, original articles submitted to Blood Transfusion, require an abstract of no more than 2,000 characters (including spaces), structured as follows: Background, Materials and methods, Results, Discussion 43 .

A good abstract should be easy to understand and broadly appealing, informative but not too detailed. It can start with a sentence or two outlining the work; then the disease and/or system studied must be introduced and what was previously unknown has to be stated in order to provide a brief overview of the current state-of-the art knowledge on the issue. The methods must be summarised without too many details; the major findings must be clearly indicated and followed by a sentence or two showing the major implications of the paper that must be consistent with the study conclusions without overestimating their possible relevance 44 . In the abstract the present tense should be used to refer to facts already established in the field, while the findings from the current study should be dealt with in the past tense.

The aim of the introduction is to introduce the topic to the readers in a straightforward way, avoiding excessive wordiness 42 . For this reason it should be short and focused, comprising approximately three paragraphs in one page 37 .

The first paragraph should mention the questions or issues that outline the background of the study and establish, using the present tense, the context, relevance, or nature of the problem, question, or purpose (what is known) 23 , 37 .

The second paragraph may include the importance of the problem and unclear issues (what is unknown).

The last paragraph should state the rationale, hypothesis, main objective, or purpose thus clearly identifying the hypothesis to be treated and the questions addressed in the manuscript (why the study was done).

One of the most common mistakes is the failure to make a clear statement of purpose. This is because many research projects, especially retrospective clinical studies, do not start at the beginning (with the identification of a specific question, followed by methods and data collection) but begin by collecting data without first identifying a specific question to be addressed that must in any case be established before beginning to write 38 . Data or conclusions from the study should not be presented or anticipated in the introduction section.

Writing the introduction at the end of the process prevents any block and it is easier after the methods, results and discussion have been completed.

Materials and methods

The methods section is one of the most important parts of a scientific manuscript and its aim is to give the reader all the necessary details to replicate the study.

CONSORT 16 , 17 , STROBE 21 , 22 and PRISMA 29 statements provide a guideline relevant to the particular type of study 2 , 42 .

The two essential elements of this section are a clear presentation of the study design and the identification and description of the measurement parameters used to evaluate the purpose of the study.

It is, therefore, necessary to provide a thorough explanation of the research methodology, including the study design, data collection, analysis principles and rationale. Special attention should be paid to the sample selection, including inclusion and exclusion criteria and to any relevant ethical considerations. A description of the randomisation or other group assignment methods used should be included, as should be the pre-specified primary and secondary outcome(s) and other variables.

According to the Uniform Requirements 9 , in the case of experimental/clinical reports involving patients or volunteers, the authors must provide information about institutional, regulatory and ethical Committee authorisation, informed consent from patients and volunteers and the observance of the latest release of the Helsinki Declaration 45 .

When reporting experiments on animals, authors should state which institutional authority granted approval for the animal experiments 9 .

Finally, in addition to describing and identifying all the measurement parameters used, it is also important to describe any unusual statistical methodology applied, how subjects were recruited and compensated and how compliance was measured (if applicable).

The results section consists of the organised presentation of the collected data. All measurements that the authors described in the materials and methods section must be reported in the results section and be presented in the same order as they were in that section 35 . The past tense should be used as results were obtained in the past. Author(s) must ensure that they use proper words when describing the relationship between data or variables. These “data relation words” should be turned into “cause/effect logic and mechanistic words” in the discussion section. A clear example of the use of this appropriate language can be found in the article by O’Connor 35 .

This section should include only data, including negative findings, and not background or methods or results of measurements that were not described in the methods section 2 . The interpretation of presented data must not be included in this section.

Results for primary and secondary outcomes can be reported using tables and figures for additional clarity. The rationale for end-point selection and the reason for the non-collection of information on important non-measured variables must be explained 35 .

Figures and tables should be simple, expand text information rather than repeat it, be consistent with reported data and summarise them 23 . In addition, they should be comprehensible on their own, that is, with only title, footnotes, abbreviations and comments.

References in this section should be limited to methods developed in the manuscript or to similar methods reported in the literature.

Patients’ anonymity is essential unless consent for publication is obtained.

The main objective of the discussion is to explain the meaning of the results.

This section should be structured as if it were a natural flow of ideas and should start with a simple statement of the key findings and whether they are consistent with the study objectives enunciated in the last paragraph of the introduction. The strengths and the limitations of the research and what the study adds to current knowledge should then be addressed 42 .

Through logical arguments, the authors should convert the relations of the variables stated in the results section into mechanistic interpretations of cause and effect using the present tense as these relations do exist at present 35 . In addition, they should describe how the results are consistent or not with similar studies and discuss any confounding factors and their impact.

They should avoid excessive wordiness and other commonly made errors such as 38 : (i) including information unrelated to the stated purpose of the article; (ii) repeating detailed data previously presented in the Results section; (iii) not interpreting and not critically analysing results of other studies reviewed and cited but rather just repeating their findings; (iv) presenting new data or new details about techniques and enrolment criteria, and (v) overstating the interpretation of the results.

Another common mistake is to forget to criticise the research described in the manuscript by highlighting the limitations of the study. The value of a scientific article is enhanced not only by showing the strengths but also the weak points of the evidence reported in the paper.

The conclusion is a separate, last paragraph that should present a concise and clear “take home” message avoiding repetition of concepts already expressed 32 . The authors should also avoid excessive generalizations of the implications of the study and remember that except for RCT there can only be testable hypotheses and observed associations, rather than rigorous proof of cause and effect 42 . Possible implications for current clinical practice or recommendations should be addressed only if appropriate.

Finally, the areas for possible improvement with future studies should be addressed avoiding ambiguous comments such as “there is a need for further research” and if there is a real need for further studies on the topic it is strongly advisable to be specific about the type of research suggested.

Acknowledgements

All contributors who do not meet the criteria for authorship should be listed in an Acknowledgements section 9 . The authors should, therefore, add a statement on the type of assistance, if any, received from the sponsor or the sponsor’s representative and include the names of any person who provided technical help, writing assistance, editorial support or any type of participation in writing the manuscript.

In addition, “when submitting a manuscript authored by a group, the corresponding author should clearly indicate the preferred citation and identify all individual authors as well as the group name. Journals generally list other members of the group in the Acknowledgments. The NLM indexes the group name and the names of individuals the group has identified as being directly responsible for the manuscript; it also lists the names of collaborators if they are listed in Acknowledgments” 9 .

The first suggestion is to follow the journal’s policies and formatting instructions, including those for books and web-based references. Other general considerations related to references, including the following ones, can be found in the Uniform Requirements 9 .

References to review articles are an efficient way to guide readers to a body of literature but they do not always reflect original work accurately. Papers accepted but not yet published should be designated as “in press” or “forthcoming” and information from manuscripts submitted but not accepted should be cited in the text as “unpublished observations”.

Avoid using abstracts as references and citing a “personal communication” unless it provides essential information not available from a public source. In this case the name of the person and date of communication should be cited in parentheses in the text. Do not include manuscripts “in submission”

In addition it is important to remember that “authors are responsible for checking that none of the references cite retracted articles except in the context of referring to the retraction. Authors can identify retracted articles in MEDLINE by using the following search term, where pt in square brackets stands for publication type: Retracted publication [pt] in PubMed” 9 . Last but not least, remember that if a reviewer does not have access to any references he or she can ask the author for a full (pdf) copy of the relevant works.

Tips for successful revision of a manuscript

Most papers are accepted after some degree of revision. In some cases, a manuscript may be rejected after internal and editorial review only.

The process of revising a manuscript and successfully responding to the comments of reviewers and Editor can be challenging. Little has been published addressing the issue of effectively revising a manuscript according to the (minor or major) comments of reviewers. This topic was recently extensively and pragmatically covered by James M. Provenzale 46 . The ten principles for revising a manuscript suggested by the author are reported in Table IV .

Ten principles for revising a manuscript suggested by James M. Provenzale 46 .

Many manuscripts are not published simply because the authors have not followed the few simple rules needed to write a good article. We hope that this paper provides the reader with the basic steps to build a draft manuscript and an outline of the process needed for publishing a manuscript. However, in Table V we summarise the ten principles we strongly recommend to comply with in order to improve the likelihood of publication of a scientific manuscript 47 .

Ten principles to improve the likelihood of publication of a scientific manuscript, suggested by James M. Provenzale 47 .

The Authors declare no conflicts of interest.

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

Research articles

scientific paper example

Educational simulator for mastoidectomy considering mechanical properties using 3D printing and its usability evaluation

  • Junhyeok Ock
  • Yeonjoo Choi

scientific paper example

A nomogram for predicting the risk of treatment failure of roxadustat in peritoneal dialysis with renal anemia

  • Jiangqing Fan

scientific paper example

Simultaneous removal of caesium and strontium using different removal mechanisms of probiotic bacteria

  • Satoshi Karasawa
  • Hideki Aoyagi

scientific paper example

Efficacy of second-line chemotherapy in patients with pulmonary large cell neuroendocrine carcinoma

  • Kazushige Wakuda
  • Toshiaki Takahashi

scientific paper example

Synthesis and biological evaluation of echinomycin analogues as potential colon cancer agent

  • Keita Kojima
  • Hiroaki Konishi
  • Satoshi Ichikawa

scientific paper example

Double-gate structure enabling remote Coulomb scattering-free transport in atomic-layer-deposited IGO thin-film transistors with HfO 2 gate dielectric through insertion of SiO 2 interlayer

  • Cheol Hee Choi
  • Jae Kyeong Jeong

scientific paper example

Clinicopathological characteristics and prognosis in patients with monoclonal gammopathy and renal damage in central China: a multicenter retrospective cohort study

  • Guolan Xing

scientific paper example

Extraction of heavy metals from copper tailings by ryegrass ( Lolium perenne L.) with the assistance of degradable chelating agents

  • Weiwei Wang
  • Jinchun Xue

scientific paper example

A methodology for image-based measurement of plate movement in disengaged wet clutches

  • Lukas Pointner-Gabriel
  • Simon Flamm
  • Karsten Stahl

scientific paper example

Competitive adsorption of CO 2 , N 2 , and CH 4 in coal-derived asphaltenes, a computational study

  • Farshad Mirzaee Valadi
  • Mohammad Pasandideh-Nadamani
  • Ayoob Torkian

scientific paper example

GPT-agents based on medical guidelines can improve the responsiveness and explainability of outcomes for traumatic brain injury rehabilitation

  • Zhang Jingfeng
  • Xia Yinshui

scientific paper example

Emission and performance analysis of diesel engine running with CeO 2 nanoparticle additive blended into castor oil biodiesel as a substitute fuel

  • Samuel Tamrat
  • Venkata Ramayya Ancha
  • Yared Seifu

scientific paper example

Mechanism of muscle atrophy in a normal-weight rat model of type 2 diabetes established by using a soft-pellet diet

  • Sayaka Akieda-Asai
  • Yukari Date

scientific paper example

Hybrid genetic algorithm-simulated annealing based electric vehicle charging station placement for optimizing distribution network resilience

  • Boya Anil Kumar
  • Milkias Berhanu Tuka

scientific paper example

Information heterogeneity between progress notes by physicians and nurses for inpatients with digestive system diseases

  • Yukinori Mashima
  • Masatoshi Tanigawa
  • Hideto Yokoi

scientific paper example

Comparison of nine trauma scoring systems in prediction of inhospital outcomes of pediatric trauma patients: a multicenter study

  • Armin Khavandegar
  • Payman Salamati

scientific paper example

Clumped isotopes reveal relationship between mussel growth and river discharge

  • Melanie A. Brewer
  • Ethan L. Grossman
  • Charles R. Randklev

scientific paper example

Analysis of post-market adverse events of istradefylline: a real-world study base on FAERS database

  • Rongrong Lu

scientific paper example

Effect of different buried depth on the disintegration characteristic of red-bed soft rock and the evolution model of disintegration breakage under cyclic drying-wetting

scientific paper example

Effect of super absorbent hydrogel on hydro-physical properties of soil under deficit irrigation

  • Rasha Abdelghafar
  • Ahmed Abdelfattah
  • Harby Mostafa

Quick links

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

scientific paper example

NASA Logo

Scientific Consensus

scientific paper example

It’s important to remember that scientists always focus on the evidence, not on opinions. Scientific evidence continues to show that human activities ( primarily the human burning of fossil fuels ) have warmed Earth’s surface and its ocean basins, which in turn have continued to impact Earth’s climate . This is based on over a century of scientific evidence forming the structural backbone of today's civilization.

NASA Global Climate Change presents the state of scientific knowledge about climate change while highlighting the role NASA plays in better understanding our home planet. This effort includes citing multiple peer-reviewed studies from research groups across the world, 1 illustrating the accuracy and consensus of research results (in this case, the scientific consensus on climate change) consistent with NASA’s scientific research portfolio.

With that said, multiple studies published in peer-reviewed scientific journals 1 show that climate-warming trends over the past century are extremely likely due to human activities. In addition, most of the leading scientific organizations worldwide have issued public statements endorsing this position. The following is a partial list of these organizations, along with links to their published statements and a selection of related resources.

American Scientific Societies

Statement on climate change from 18 scientific associations.

"Observations throughout the world make it clear that climate change is occurring, and rigorous scientific research demonstrates that the greenhouse gases emitted by human activities are the primary driver." (2009) 2

American Association for the Advancement of Science

"Based on well-established evidence, about 97% of climate scientists have concluded that human-caused climate change is happening." (2014) 3

AAAS emblem

American Chemical Society

"The Earth’s climate is changing in response to increasing concentrations of greenhouse gases (GHGs) and particulate matter in the atmosphere, largely as the result of human activities." (2016-2019) 4

ACS emblem

American Geophysical Union

"Based on extensive scientific evidence, it is extremely likely that human activities, especially emissions of greenhouse gases, are the dominant cause of the observed warming since the mid-20th century. There is no alterative explanation supported by convincing evidence." (2019) 5

AGU emblem

American Medical Association

"Our AMA ... supports the findings of the Intergovernmental Panel on Climate Change’s fourth assessment report and concurs with the scientific consensus that the Earth is undergoing adverse global climate change and that anthropogenic contributions are significant." (2019) 6

AMA emblem

American Meteorological Society

"Research has found a human influence on the climate of the past several decades ... The IPCC (2013), USGCRP (2017), and USGCRP (2018) indicate that it is extremely likely that human influence has been the dominant cause of the observed warming since the mid-twentieth century." (2019) 7

AMS emblem

American Physical Society

"Earth's changing climate is a critical issue and poses the risk of significant environmental, social and economic disruptions around the globe. While natural sources of climate variability are significant, multiple lines of evidence indicate that human influences have had an increasingly dominant effect on global climate warming observed since the mid-twentieth century." (2015) 8

APS emblem

The Geological Society of America

"The Geological Society of America (GSA) concurs with assessments by the National Academies of Science (2005), the National Research Council (2011), the Intergovernmental Panel on Climate Change (IPCC, 2013) and the U.S. Global Change Research Program (Melillo et al., 2014) that global climate has warmed in response to increasing concentrations of carbon dioxide (CO2) and other greenhouse gases ... Human activities (mainly greenhouse-gas emissions) are the dominant cause of the rapid warming since the middle 1900s (IPCC, 2013)." (2015) 9

GSA emblem

Science Academies

International academies: joint statement.

"Climate change is real. There will always be uncertainty in understanding a system as complex as the world’s climate. However there is now strong evidence that significant global warming is occurring. The evidence comes from direct measurements of rising surface air temperatures and subsurface ocean temperatures and from phenomena such as increases in average global sea levels, retreating glaciers, and changes to many physical and biological systems. It is likely that most of the warming in recent decades can be attributed to human activities (IPCC 2001)." (2005, 11 international science academies) 1 0

U.S. National Academy of Sciences

"Scientists have known for some time, from multiple lines of evidence, that humans are changing Earth’s climate, primarily through greenhouse gas emissions." 1 1

UNSAS emblem

U.S. Government Agencies

U.s. global change research program.

"Earth’s climate is now changing faster than at any point in the history of modern civilization, primarily as a result of human activities." (2018, 13 U.S. government departments and agencies) 12

USGCRP emblem

Intergovernmental Bodies

Intergovernmental panel on climate change.

“It is unequivocal that the increase of CO 2 , methane, and nitrous oxide in the atmosphere over the industrial era is the result of human activities and that human influence is the principal driver of many changes observed across the atmosphere, ocean, cryosphere, and biosphere. “Since systematic scientific assessments began in the 1970s, the influence of human activity on the warming of the climate system has evolved from theory to established fact.” 1 3-17

IPCC emblem

Other Resources

List of worldwide scientific organizations.

The following page lists the nearly 200 worldwide scientific organizations that hold the position that climate change has been caused by human action. http://www.opr.ca.gov/facts/list-of-scientific-organizations.html

U.S. Agencies

The following page contains information on what federal agencies are doing to adapt to climate change. https://www.c2es.org/site/assets/uploads/2012/02/climate-change-adaptation-what-federal-agencies-are-doing.pdf

Technically, a “consensus” is a general agreement of opinion, but the scientific method steers us away from this to an objective framework. In science, facts or observations are explained by a hypothesis (a statement of a possible explanation for some natural phenomenon), which can then be tested and retested until it is refuted (or disproved).

As scientists gather more observations, they will build off one explanation and add details to complete the picture. Eventually, a group of hypotheses might be integrated and generalized into a scientific theory, a scientifically acceptable general principle or body of principles offered to explain phenomena.

1. K. Myers, et al, "Consensus revisited: quantifying scientific agreement on climate change and climate expertise among Earth scientists 10 years later", Environmental Research Letters Vol.16 No. 10, 104030 (20 October 2021); DOI:10.1088/1748-9326/ac2774 M. Lynas, et al, "Greater than 99% consensus on human caused climate change in the peer-reviewed scientific literature", Environmental Research Letters Vol.16 No. 11, 114005 (19 October 2021); DOI:10.1088/1748-9326/ac2966 J. Cook et al., "Consensus on consensus: a synthesis of consensus estimates on human-caused global warming", Environmental Research Letters Vol. 11 No. 4, (13 April 2016); DOI:10.1088/1748-9326/11/4/048002 J. Cook et al., "Quantifying the consensus on anthropogenic global warming in the scientific literature", Environmental Research Letters Vol. 8 No. 2, (15 May 2013); DOI:10.1088/1748-9326/8/2/024024 W. R. L. Anderegg, “Expert Credibility in Climate Change”, Proceedings of the National Academy of Sciences Vol. 107 No. 27, 12107-12109 (21 June 2010); DOI: 10.1073/pnas.1003187107 P. T. Doran & M. K. Zimmerman, "Examining the Scientific Consensus on Climate Change", Eos Transactions American Geophysical Union Vol. 90 Issue 3 (2009), 22; DOI: 10.1029/2009EO030002 N. Oreskes, “Beyond the Ivory Tower: The Scientific Consensus on Climate Change”, Science Vol. 306 no. 5702, p. 1686 (3 December 2004); DOI: 10.1126/science.1103618

2. Statement on climate change from 18 scientific associations (2009)

3. AAAS Board Statement on Climate Change (2014)

4. ACS Public Policy Statement: Climate Change (2016-2019)

5. Society Must Address the Growing Climate Crisis Now (2019)

6. Global Climate Change and Human Health (2019)

7. Climate Change: An Information Statement of the American Meteorological Society (2019)

8. American Physical Society (2021)

9. GSA Position Statement on Climate Change (2015)

10. Joint science academies' statement: Global response to climate change (2005)

11. Climate at the National Academies

12. Fourth National Climate Assessment: Volume II (2018)

13. IPCC Fifth Assessment Report, Summary for Policymakers, SPM 1.1 (2014)

14. IPCC Fifth Assessment Report, Summary for Policymakers, SPM 1 (2014)

15. IPCC Sixth Assessment Report, Working Group 1 (2021)

16. IPCC Sixth Assessment Report, Working Group 2 (2022)

17. IPCC Sixth Assessment Report, Working Group 3 (2022)

Discover More Topics From NASA

Explore Earth Science

scientific paper example

Earth Science in Action

Earth Action

Earth Science Data

The sum of Earth's plants, on land and in the ocean, changes slightly from year to year as weather patterns shift.

Facts About Earth

scientific paper example

Help | Advanced Search

Computer Science > Computation and Language

Title: can llms learn from previous mistakes investigating llms' errors to boost for reasoning.

Abstract: Recent works have shown the benefits to LLMs from fine-tuning golden-standard Chain-of-Thought (CoT) rationales or using them as correct examples in few-shot prompting. While humans can indeed imitate correct examples, learning from our mistakes is another vital aspect of human cognition. Hence, a question naturally arises: \textit{can LLMs learn and benefit from their mistakes, especially for their reasoning? } This study investigates this problem from both the prompting and model-tuning perspectives. We begin by introducing \textsc{CoTErrorSet}, a new benchmark with 609,432 questions, each designed with both correct and error references, and demonstrating the types and reasons for making such mistakes. To explore the effectiveness of those mistakes, we design two methods: (1) \textbf{Self-rethinking} prompting guides LLMs to rethink whether they have made similar previous mistakes; and (2) \textbf{Mistake tuning} involves finetuning models in both correct and incorrect reasoning domains, rather than only tuning models to learn ground truth in traditional methodology. We conduct a series of experiments to prove LLMs can obtain benefits from mistakes in both directions. Our two methods offer potentially cost-effective strategies by leveraging errors to enhance reasoning capabilities, which costs significantly less than creating meticulously hand-crafted golden references. We ultimately make a thorough analysis of the reasons behind LLMs' errors, which provides directions that future research needs to overcome. \textsc{CoTErrorSet} will be published soon on \texttt{Anonymity Link}.

Submission history

Access paper:.

  • HTML (experimental)
  • Other Formats

References & Citations

  • Google Scholar
  • Semantic Scholar

BibTeX formatted citation

BibSonomy logo

Bibliographic and Citation Tools

Code, data and media associated with this article, recommenders and search tools.

  • Institution

arXivLabs: experimental projects with community collaborators

arXivLabs is a framework that allows collaborators to develop and share new arXiv features directly on our website.

Both individuals and organizations that work with arXivLabs have embraced and accepted our values of openness, community, excellence, and user data privacy. arXiv is committed to these values and only works with partners that adhere to them.

Have an idea for a project that will add value for arXiv's community? Learn more about arXivLabs .

IMAGES

  1. Examples Of Science Paper Abstract / Research Paper Sample Pdf Chapter

    scientific paper example

  2. Scientific Research Paper Sample

    scientific paper example

  3. Format Example Of Scientific Paper / Scientific journal article example

    scientific paper example

  4. Research Paper Format

    scientific paper example

  5. How to Write a Scientific Paper

    scientific paper example

  6. Examples Of Science Paper Abstract / Sample project abstract

    scientific paper example

VIDEO

  1. Scientific Paper Presentation Session 15 Academic English

  2. FOR A KILLER ABSTRACT

  3. Structuring Scientific Paper

  4. Lecture 1: Foundations of Knowledge, Science, Research and Scientific/Academic Publishing

  5. How to Write a Scientific Research Paper

  6. How to write a scientific paper: Paper structure #academia #graduateschool #research #graduates

COMMENTS

  1. Sample Paper in Scientific Format

    Learn how to write a scientific paper with this example from a biology course. See the structure, format, and content of the title, introduction, methods, results, and discussion sections.

  2. The Scientific Paper: A Template

    This document provides an outline and tips for writing scientific papers in MS Word. It includes sections on abstract, introduction, methods, results, discussion, and references, with examples and formatting guidelines.

  3. How to Format a Scientific Paper

    Learn the rules and guidelines for formatting a scientific paper according to the IMRaD structure. Find out how to write the cover page, abstract, keywords, introduction, methods, results, discussion, conclusion, and references.

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

    While articles in the journal Science are especially concise and lack the divisions of a normal scientific paper, Kilner et al. offers plenty of examples of effective communication strategies that are utilized in scientific writing. We hope that the guidelines that follow, as well as the concrete examples provided, will lead to scientific ...

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

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

  6. Scientific Papers

    Learn how to write and structure scientific papers that inform, convince, and persuade your audience. Find out the key components of an effective introduction, such as context, need, task, and object of the document.

  7. 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. ... For example, when a study is conducted on humans, you must provide, up front, assurance that it has received the approval of you Institution Ethics ...

  8. HOW TO WRITE A SCIENTIFIC ARTICLE

    The task of writing a scientific paper and submitting it to a journal for publication is a time‐consuming and often daunting task. 3,4 Barriers to effective writing include lack ... It may be helpful to follow the IMRaD format for writing scientific manuscripts. This acronym stands for the sections contained within the article: Introduction ...

  9. Formatting guide

    A useful set of articles providing general advice about writing and submitting scientific papers can be found on ... that summarizes the main finding of the paper, where appropriate (for example ...

  10. How to write a first-class paper

    In each paragraph, the first sentence defines the context, the body contains the new idea and the final sentence offers a conclusion. For the whole paper, the introduction sets the context, the ...

  11. Writing a scientific article: A step-by-step guide for beginners

    The vast majority of scientific journals follow the so-called "IMRAD" format, i.e. introduction, methods, results and discussion. Naturally, there are some exceptions to this rule, and you should always check the instructions for authors of the journal where you plan to submit your paper to ensure that this is indeed the recommended format.

  12. Writing a Research Paper Introduction

    In a paper structured using the standard scientific "introduction, methods, results, discussion" format, this isn't always necessary. But if your paper is structured in a less predictable way, it's important to describe the shape of it for the reader. If included, the overview should be concise, direct, and written in the present tense.

  13. Guide: Writing the Scientific Paper

    Learn how to write a scientific paper with a standard format of six main parts: title, abstract, introduction, methods and materials, results, and discussion. See examples of scientific papers from different fields and learn how to draft them effectively.

  14. PDF Writing a scientific paper, step by painful step

    To polish this turd takes more than checking for typos; there are four critical steps. (1) read and revise the paper until you believe it is complete and coherent. (2) confirm the paragraph and sentence structure/flow. (3) correct common wording mistakes. (4) proof it and send the draft out for comments.

  15. Research Guides: Writing a Scientific Paper: INTRODUCTION

    Chris A. Mack. SPIE. 2018. Indicate the field of the work, why this field is important, and what has already been done (with proper citations). Indicate a gap, raise a research question, or challenge prior work in this territory. Outline the purpose and announce the present research, clearly indicating what is novel and why it is significant.

  16. How do I write a scientific paper?

    Learn the basic structure and format of a scientific paper, from title and abstract to results and discussion. Find tips on how to present your research results, avoid jargon and acronyms, and cite references.

  17. Scientific Reports

    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.

  18. Instructions for preparing an initial manuscript

    2. A. Z. Author, "Example conference paper" in Proceedings of the 2021 Example Conference on Interesting Topic (Publisher, 2021), abstr. 875. [published, paper title and abstract number provided] 3. M. Author, paper presented at the 25th Annual Meeting of Example Organization, Washington, DC, 10 to 13 October 2022. [unpublished] Theses

  19. How to write a scientific manuscript for publication

    The anatomy of a paper: from origin to current format. The history of scientific journals dates from 1665, when the French "Journal des sçavans" and the English "Philosophical Transactions of the Royal Society" first began systematically publishing research results 7.From then on, the initial structure of scientific papers evolved gradually from letters (usually by a single author ...

  20. PDF The Structure of an Academic Paper

    example, although the hourglass model introduces the most important point right from the beginning as a guide to the rest of the paper, some traditions build the argument gradually and deliver the main idea as a punchline. Scholars in the U.S. sometimes joke that the style is to "tell your audience what you are going to tell them,

  21. PDF Sample Scientific Paper

    Sample Scientific Paper. Method and Materials To test our hypothesis, we used the chick embryo model. Mouse mammary tumor cells (4T1) and murine macrophages (J774A.1) or murine adipocytes (3T3-L1) were inoculated on the chick chorioallantoic membrane (CAM) on the 10th embryonic day. The primary tumor and the chick

  22. The 100 most-cited scientific papers

    Here at Science we love ranking things, so we were thrilled with this list of the top 100 most-cited scientific papers, courtesy of Nature.Surprisingly absent are many of the landmark discoveries you might expect, such as the discovery of DNA's double helix structure. Instead, most of these influential manuscripts are slightly more utilitarian in nature.

  23. Research articles

    Read the latest Research articles from Scientific Reports. ... Calls for Papers Guide to referees Editor's Choice Journal highlights Publish with us ...

  24. Generative emulation of weather forecast ensembles with ...

    Weather is inherently uncertain, making the assessment of forecast uncertainty a vital component of operational weather forecasting (1, 2).Given a numerical weather prediction (NWP) model, the standard way to quantify this uncertainty is to stochastically perturb the model's initial conditions and its representation of small-scale physical processes to create an ensemble of weather ...

  25. Scientific Consensus

    Technically, a "consensus" is a general agreement of opinion, but the scientific method steers us away from this to an objective framework. In science, facts or observations are explained by a hypothesis (a statement of a possible explanation for some natural phenomenon), which can then be tested and retested until it is refuted (or disproved).

  26. [2403.20046] Can LLMs Learn from Previous Mistakes? Investigating LLMs

    Recent works have shown the benefits to LLMs from fine-tuning golden-standard Chain-of-Thought (CoT) rationales or using them as correct examples in few-shot prompting. While humans can indeed imitate correct examples, learning from our mistakes is another vital aspect of human cognition. Hence, a question naturally arises: \\textit{can LLMs learn and benefit from their mistakes, especially ...