Scientific and Scholarly Writing

  • Literature Searches
  • Tracking and Citing References

Parts of a Scientific & Scholarly Paper

Introduction.

  • Writing Effectively
  • Where to Publish?
  • Capstone Resources

Different sections are needed in different types of scientific papers (lab reports, literature reviews, systematic reviews, methods papers, research papers, etc.). Projects that overlap with the social sciences or humanities may have different requirements. Generally, however, you'll need to include:

INTRODUCTION (Background)

METHODS SECTION (Materials and Methods)

What is a title

Titles have two functions: to identify the main topic or the message of the paper and to attract readers.

The title will be read by many people. Only a few will read the entire paper, therefore all words in the title should be chosen with care. Too short a title is not helpful to the potential reader. Too long a title can sometimes be even less meaningful. Remember a title is not an abstract. Neither is a title a sentence.

What makes a good title?

A good title is accurate, complete, and specific. Imagine searching for your paper in PubMed. What words would you use?

  • Use the fewest possible words that describe the contents of the paper.
  • Avoid waste words like "Studies on", or "Investigations on".
  • Use specific terms rather than general.
  • Use the same key terms in the title as the paper.
  • Watch your word order and syntax.

The abstract is a miniature version of your paper. It should present the main story and a few essential details of the paper for readers who only look at the abstract and should serve as a clear preview for readers who read your whole paper. They are usually short (250 words or less).

The goal is to communicate:

  •  What was done?
  •  Why was it done?
  •  How was it done?
  •  What was found?

A good abstract is specific and selective. Try summarizing each of the sections of your paper in a sentence two. Do the abstract last, so you know exactly what you want to write.

  • Use 1 or more well developed paragraphs.
  • Use introduction/body/conclusion structure.
  • Present purpose, results, conclusions and recommendations in that order.
  • Make it understandable to a wide audience.
  • << Previous: Tracking and Citing References
  • Next: Writing Effectively >>
  • Last Updated: Apr 17, 2024 3:27 PM
  • URL: https://libraryguides.umassmed.edu/scientific-writing

Home

  • Duke NetID Login
  • 919.660.1100
  • Duke Health Badge: 24-hour access
  • Accounts & Access
  • Databases, Journals & Books
  • Request & Reserve
  • Training & Consulting
  • Request Articles & Books
  • Renew Online
  • Reserve Spaces
  • Reserve a Locker
  • Study & Meeting Rooms
  • Course Reserves
  • Digital Health Device Collection
  • Pay Fines/Fees
  • Recommend a Purchase
  • Access From Off Campus
  • Building Access
  • Computers & Equipment
  • Wifi Access
  • My Accounts
  • Mobile Apps
  • Known Access Issues
  • Report an Access Issue
  • All Databases
  • Article Databases
  • Basic Sciences
  • Clinical Sciences
  • Dissertations & Theses
  • Drugs, Chemicals & Toxicology
  • Grants & Funding
  • Interprofessional Education
  • Non-Medical Databases
  • Search for E-Journals
  • Search for Print & E-Journals
  • Search for E-Books
  • Search for Print & E-Books
  • E-Book Collections
  • Biostatistics
  • Global Health
  • MBS Program
  • Medical Students
  • MMCi Program
  • Occupational Therapy
  • Path Asst Program
  • Physical Therapy
  • Researchers
  • Community Partners

Conducting Research

  • Archival & Historical Research
  • Black History at Duke Health
  • Data Analytics & Viz Software
  • Data: Find and Share
  • Evidence-Based Practice
  • NIH Public Access Policy Compliance
  • Publication Metrics
  • Qualitative Research
  • Searching Animal Alternatives
  • Systematic Reviews
  • Test Instruments

Using Databases

  • JCR Impact Factors
  • Web of Science

Finding & Accessing

  • COVID-19: Core Clinical Resources
  • Health Literacy
  • Health Statistics & Data
  • Library Orientation

Writing & Citing

  • Creating Links
  • Getting Published
  • Reference Mgmt
  • Scientific Writing

Meet a Librarian

  • Request a Consultation
  • Find Your Liaisons
  • Register for a Class
  • Request a Class
  • Self-Paced Learning

Search Services

  • Literature Search
  • Systematic Review
  • Animal Alternatives (IACUC)
  • Research Impact

Citation Mgmt

  • Other Software

Scholarly Communications

  • About Scholarly Communications
  • Publish Your Work
  • Measure Your Research Impact
  • Engage in Open Science
  • Libraries and Publishers
  • Directions & Maps
  • Floor Plans

Library Updates

  • Annual Snapshot
  • Conference Presentations
  • Contact Information
  • Gifts & Donations

Scientific Writing: Sections of a Paper

  • Sections of a Paper
  • Common Grammar Mistakes Explained
  • Citing Sources

Introduction

  • Materials & Methods

Typically scientific journal articles have the following sections:

Materials & Methods

References used:

Kotsis, S.V. and Chung, K.C. (2010) A Guide for Writing in the Scientific Forum. Plastic and Reconstructive Surgery. 126(5):1763-71. PubMed ID:  21042135

Van Way, C.W. (2007) Writing a Scientific Paper. Nutrition in Clinical Practice. 22: 663-40. PubMed ID:  1804295

What to include:

  • Background/Objectives: include the hypothesis
  • Methods: Briefly explain the type of study, sample/population size and description, the design, and any particular techniques for data collection and analysis
  • Results: Essential data, including statistically significant data (use # & %)
  • Conclusions: Summarize interpretations of results and explain if hypothesis was supported or rejected
  • Be concise!
  • Emphasize the methods and results
  • Do not copy the introduction
  • Only include data that is included in the paper
  • Write the abstract last
  • Avoid jargon and ambiguity
  • Should stand-alone

Additional resources: Fisher, W. E. (2005) Abstract Writing. Journal of Surgical Research. 128(2):162-4. PubMed ID:  16165161 Peh, W.C. and Ng, K.H. (2008) Abstract and keywords. Singapore Medical Journal. 49(9): 664-6. PubMed ID:  18830537

  • How does your study fit into what has been done
  • Explain evidence using limited # of references
  • Why is it important
  • How does it relate to previous research
  • State hypothesis at the end
  • Use present tense
  • Be succinct
  • Clearly state objectives
  • Explain important work done

Additional resources: Annesley, T. M. (2010) "It was a cold and rainy night": set the scene with a good introduction. Clinical Chemistry. 56(5):708-13. PubMed ID:  20207764 Peh, W.C. and Ng, K.H. (2008) Writing the introduction. Singapore Medical Journal. 49(10):756-8. PubMed ID:  18946606  

  • What was done
  • Include characteristics
  • Describe recruitment, participation, withdrawal, etc.
  • Type of study (RCT, cohort, case-controlled, etc.)
  • Equipment used
  • Measurements made
  • Usually the final paragraph
  • Include enough details so others can duplicate study
  • Use past tense
  • Be direct and precise
  • Include any preliminary results
  • Ask for help from a statistician to write description of statistical analysis
  • Be systematic

Additional resources: Lallet, R. H. (2004) How to write the methods section of a research paper. Respiratory Care. 49(10): 1229-32. PubMed ID:  15447808 Ng, K.H. and Peh, W.C. (2008) Writing the materials and methods. Singapore Medical Journal. 49(11): 856-9. PubMed ID:  19037549

  • Describe study sample demographics
  • Include statistical significance and the statistical test used
  • Use tables and figures when appropriate
  • Present in a logical sequence
  • Facts only - no citations or interpretations
  • Should stand alone (not need written descriptions to be understood)
  • Include title, legend, and axes labels
  • Include raw numbers with percentages
  • General phrases (significance, show trend, etc. should be used with caution)
  • Data is plural ("Our data are" is correct, "Our data is" is in-correct)

Additional resources: Ng, K.H and Peh, W.C. (2008) Writing the results. Singapore Medical Journal. 49(12):967-9. PubMed ID:  19122944 Streiner, D.L. (2007) A shortcut to rejection: how not to write the results section of a paper. Canadian Journal of Psychiatry. 52(6):385-9. PubMed ID:  17696025

  • Did you reject your null hypothesis?
  • Include a focused review of literature in relation to results
  • Explain meaning of statistical findings
  • Explain importance/relevance 
  • Include all possible explanations
  • Discuss possible limitations of study
  • Suggest future work that could be done
  • Use past tense to describe your study and present tense to describe established knowledge from literature
  • Don't criticize other studies, contrast it with your work
  • Don't make conclusions not supported by your results
  • Stay focused and concise
  • Include key, relevant references
  • It is considered good manners to include an acknowledgements section

Additional resources: Annesley, T. M. (2010) The discussion section: your closing argument. Clinical Chemistry. 56(11):1671-4. PubMed ID:  20833779 Ng, K.H. and Peh, W.C. (2009) Writing the discussion. Singapore Medical Journal. 50(5):458-61. PubMed ID:  19495512

Tables & Figures: Durbin, C. G. (2004) Effective use of tables and figures in abstracts, presentations, and papers. Respiratory Care. 49(10): 1233-7. PubMed ID:  15447809 Ng, K. H. and Peh, W.C.G. (2009) Preparing effective tables. Singapore Medical Journal. (50)2: 117-9. PubMed ID:  19296024

Statistics: Ng, K. H. and Peh, W.C.G. (2009) Presenting the statistical results. Singapore Medical Journal. (50)1: 11-4. PubMed ID:  19224078

References: Peh, W.C.G. and Ng, K. H. (2009) Preparing the references. Singapore Medical Journal. (50)7: 11-4. PubMed ID:  19644619

Additional Resources

  • More from Elsevier Elsevier's Research Academy is an online tutorial to help with writing books, journals, and grants. It also includes information on citing sources, peer reviewing, and ethics in publishing
  • Research4Life Training Portal Research4Life provides downloadable instruction materials, including modules on authorship skills as well as other research related skills.
  • Coursera: Science Writing Coursera provides a wide variety of online courses for continuing education. You can search around for various courses on scientific writing or academic writing, and they're available to audit for free.

parts of a science research paper

  • << Previous: Lit Review
  • Next: Grammar/Language >>
  • Last Updated: Mar 20, 2024 2:34 PM
  • URL: https://guides.mclibrary.duke.edu/scientificwriting
  • Duke Health
  • Duke University
  • Duke Libraries
  • Medical Center Archives
  • Duke Directory
  • Seeley G. Mudd Building
  • 10 Searle Drive
  • [email protected]

This page has been archived and is no longer updated

Scientific Papers

Scientific papers are for sharing your own original research work with other scientists or for reviewing the research conducted by others. As such, they are critical to the evolution of modern science, in which the work of one scientist builds upon that of others. To reach their goal, papers must aim to inform, not impress. They must be highly readable — that is, clear, accurate, and concise. They are more likely to be cited by other scientists if they are helpful rather than cryptic or self-centered.

Scientific papers typically have two audiences: first, the referees, who help the journal editor decide whether a paper is suitable for publication; and second, the journal readers themselves, who may be more or less knowledgeable about the topic addressed in the paper. To be accepted by referees and cited by readers, papers must do more than simply present a chronological account of the research work. Rather, they must convince their audience that the research presented is important, valid, and relevant to other scientists in the same field. To this end, they must emphasize both the motivation for the work and the outcome of it, and they must include just enough evidence to establish the validity of this outcome.

Papers that report experimental work are often structured chronologically in five sections: first, Introduction ; then Materials and Methods , Results , and Discussion (together, these three sections make up the paper's body); and finally, Conclusion .

  • The Introduction section clarifies the motivation for the work presented and prepares readers for the structure of the paper.
  • The Materials and Methods section provides sufficient detail for other scientists to reproduce the experiments presented in the paper. In some journals, this information is placed in an appendix, because it is not what most readers want to know first.
  • The Results and Discussion sections present and discuss the research results, respectively. They are often usefully combined into one section, however, because readers can seldom make sense of results alone without accompanying interpretation — they need to be told what the results mean.
  • The Conclusion section presents the outcome of the work by interpreting the findings at a higher level of abstraction than the Discussion and by relating these findings to the motivation stated in the Introduction .

(Papers reporting something other than experiments, such as a new method or technology, typically have different sections in their body, but they include the same Introduction and Conclusion sections as described above.)

Although the above structure reflects the progression of most research projects, effective papers typically break the chronology in at least three ways to present their content in the order in which the audience will most likely want to read it. First and foremost, they summarize the motivation for, and the outcome of, the work in an abstract, located before the Introduction . In a sense, they reveal the beginning and end of the story — briefly — before providing the full story. Second, they move the more detailed, less important parts of the body to the end of the paper in one or more appendices so that these parts do not stand in the readers' way. Finally, they structure the content in the body in theorem-proof fashion, stating first what readers must remember (for example, as the first sentence of a paragraph) and then presenting evidence to support this statement.

The introduction

  • First, provide some context to orient those readers who are less familiar with your topic and to establish the importance of your work.
  • Second, state the need for your work, as an opposition between what the scientific community currently has and what it wants.
  • Third, indicate what you have done in an effort to address the need (this is the task).
  • Finally, preview the remainder of the paper to mentally prepare readers for its structure, in the object of the document.

Context and need

At the beginning of the Introduction section, the context and need work together as a funnel: They start broad and progressively narrow down to the issue addressed in the paper. To spark interest among your audience — referees and journal readers alike — provide a compelling motivation for the work presented in your paper: The fact that a phenomenon has never been studied before is not, in and of itself, a reason to study that phenomenon.

Write the context in a way that appeals to a broad range of readers and leads into the need. Do not include context for the sake of including context: Rather, provide only what will help readers better understand the need and, especially, its importance. Consider anchoring the context in time, using phrases such as recently , in the past 10 years , or since the early 1990s . You may also want to anchor your context in space (either geographically or within a given research field).

Convey the need for the work as an opposition between actual and desired situations. Start by stating the actual situation (what we have) as a direct continuation of the context. If you feel you must explain recent achievements in much detail — say, in more than one or two paragraphs — consider moving the details to a section titled State of the art (or something similar) after the Introduction , but do provide a brief idea of the actual situation in the Introduction . Next, state the desired situation (what we want). Emphasize the contrast between the actual and desired situations with such words as but , however, or unfortunately .

One elegant way to express the desired part of the need is to combine it with the task in a single sentence. This sentence expresses first the objective, then the action undertaken to reach this objective, thus creating a strong and elegant connection between need and task. Here are three examples of such a combination:

To confirm this assumption , we studied the effects of a range of inhibitors of connexin channels . . . on . . .
To assess whether such multiple-coil sensors perform better than single-signal ones , we tested two of them — the DuoPXK and the GEMM3 — in a field where . . . To form a better view of the global distribution and infectiousness of this pathogen , we examined 1645 postmetamorphic and adult amphibians collected from 27 countries between 1984 and 2006 for the presence of . . .

Task and object

An Introduction is usually clearer and more logical when it separates what the authors have done (the task) from what the paper itself attempts or covers (the object of the document). In other words, the task clarifies your contribution as a scientist, whereas the object of the document prepares readers for the structure of the paper, thus allowing focused or selective reading.

For the task,

  • use whoever did the work (normally, you and your colleagues) as the subject of the sentence: we or perhaps the authors;
  • use a verb expressing a research action: measured , calculated , etc.;
  • set that verb in the past tense.

The three examples below are well-formed tasks.

To confirm this assumption, we studied the effects of a range of inhibitors of connexin channels, such as the connexin mimetic peptides Gap26 and Gap27 and anti-peptide antibodies, on calcium signaling in cardiac cells and HeLa cells expressing connexins.
During controlled experiments, we investigated the influence of the HMP boundary conditions on liver flows.
To tackle this problem, we developed a new software verification technique called oblivious hashing, which calculates the hash values based on the actual execution of the program.

The list below provides examples of verbs that express research actions:

For the object of the document,

  • use the document itself as the subject of the sentence: this paper , this letter , etc.;
  • use a verb expressing a communication action: presents , summarizes , etc.;
  • set the verb in the present tense.

The three examples below are suitable objects of the document for the three tasks shown above, respectively.

This paper clarifies the role of CxHc on calcium oscillations in neonatal cardiac myocytes and calcium transients induced by ATP in HL-cells originated from cardiac atrium and in HeLa cells expressing connexin 43 or 26. This paper presents the flow effects induced by increasing the hepatic-artery pressure and by obstructing the vena cava inferior. This paper discusses the theory behind oblivious hashing and shows how this approach can be applied for local software tamper resistance and remote code authentication.

The list below provides examples of verbs that express communication actions:

Even the most logical structure is of little use if readers do not see and understand it as they progress through a paper. Thus, as you organize the body of your paper into sections and perhaps subsections, remember to prepare your readers for the structure ahead at all levels. You already do so for the overall structure of the body (the sections) in the object of the document at the end of the Introduction . You can similarly prepare your readers for an upcoming division into subsections by introducing a global paragraph between the heading of a section and the heading of its first subsection. This paragraph can contain any information relating to the section as a whole rather than particular subsections, but it should at least announce the subsections, whether explicitly or implicitly. An explicit preview would be phrased much like the object of the document: "This section first . . . , then . . . , and finally . . . "

Although papers can be organized into sections in many ways, those reporting experimental work typically include Materials and Methods , Results , and Discussion in their body. In any case, the paragraphs in these sections should begin with a topic sentence to prepare readers for their contents, allow selective reading, and — ideally — get a message across.

Materials and methods

Results and discussion.

When reporting and discussing your results, do not force your readers to go through everything you went through in chronological order. Instead, state the message of each paragraph upfront: Convey in the first sentence what you want readers to remember from the paragraph as a whole. Focus on what happened, not on the fact that you observed it. Then develop your message in the remainder of the paragraph, including only that information you think you need to convince your audience.

The conclusion

At the end of your Conclusion , consider including perspectives — that is, an idea of what could or should still be done in relation to the issue addressed in the paper. If you include perspectives, clarify whether you are referring to firm plans for yourself and your colleagues ("In the coming months, we will . . . ") or to an invitation to readers ("One remaining question is . . . ").

If your paper includes a well-structured Introduction and an effective abstract, you need not repeat any of the Introduction in the Conclusion . In particular, do not restate what you have done or what the paper does. Instead, focus on what you have found and, especially, on what your findings mean. Do not be afraid to write a short Conclusion section: If you can conclude in just a few sentences given the rich discussion in the body of the paper, then do so. (In other words, resist the temptation to repeat material from the Introduction just to make the Conclusio n longer under the false belief that a longer Conclusion will seem more impressive.)

The abstract

Typically, readers are primarily interested in the information presented in a paper's Introduction and Conclusion sections. Primarily, they want to know the motivation for the work presented and the outcome of this work. Then (and only then) the most specialized among them might want to know the details of the work. Thus, an effective abstract focuses on motivation and outcome; in doing so, it parallels the paper's Introduction and Conclusion .

Accordingly, you can think of an abstract as having two distinct parts — motivation and outcome — even if it is typeset as a single paragraph. For the first part, follow the same structure as the Introduction section of the paper: State the context, the need, the task, and the object of the document. For the second part, mention your findings (the what ) and, especially, your conclusion (the so what — that is, the interpretation of your findings); if appropriate, end with perspectives, as in the Conclusion section of your paper.

Although the structure of the abstract parallels the Introduction and Conclusion sections, it differs from these sections in the audience it addresses. The abstract is read by many different readers, from the most specialized to the least specialized among the target audience. In a sense, it should be the least specialized part of the paper. Any scientist reading it should be able to understand why the work was carried out and why it is important (context and need), what the authors did (task) and what the paper reports about this work (object of the document), what the authors found (findings), what these findings mean (the conclusion), and possibly what the next steps are (perspectives). In contrast, the full paper is typically read by specialists only; its Introduction and Conclusion are more detailed (that is, longer and more specialized) than the abstract.

An effective abstract stands on its own — it can be understood fully even when made available without the full paper. To this end, avoid referring to figures or the bibliography in the abstract. Also, introduce any acronyms the first time you use them in the abstract (if needed), and do so again in the full paper (see Mechanics: Using abbreviations ).

This page appears in the following eBook

Topic rooms within Scientific Communication

Topic Rooms

Within this Subject (22)

  • Communicating as a Scientist (3)
  • Papers (4)
  • Correspondence (5)
  • Presentations (4)
  • Conferences (3)
  • Classrooms (3)

Other Topic Rooms

  • Gene Inheritance and Transmission
  • Gene Expression and Regulation
  • Nucleic Acid Structure and Function
  • Chromosomes and Cytogenetics
  • Evolutionary Genetics
  • Population and Quantitative Genetics
  • Genes and Disease
  • Genetics and Society
  • Cell Origins and Metabolism
  • Proteins and Gene Expression
  • Subcellular Compartments
  • Cell Communication
  • Cell Cycle and Cell Division

ScholarCast

© 2014 Nature Education

  • Press Room |
  • Terms of Use |
  • Privacy Notice |

Send

Visual Browse

  • Research Guides

BSCI 1510L Literature and Stats Guide: 3.2 Components of a scientific paper

  • 1 What is a scientific paper?
  • 2 Referencing and accessing papers
  • 2.1 Literature Cited
  • 2.2 Accessing Scientific Papers
  • 2.3 Traversing the web of citations
  • 2.4 Keyword Searches
  • 3 Style of scientific writing
  • 3.1 Specific details regarding scientific writing

3.2 Components of a scientific paper

  • 4 For further information
  • Appendix A: Calculation Final Concentrations
  • 1 Formulas in Excel
  • 2 Basic operations in Excel
  • 3 Measurement and Variation
  • 3.1 Describing Quantities and Their Variation
  • 3.2 Samples Versus Populations
  • 3.3 Calculating Descriptive Statistics using Excel
  • 4 Variation and differences
  • 5 Differences in Experimental Science
  • 5.1 Aside: Commuting to Nashville
  • 5.2 P and Detecting Differences in Variable Quantities
  • 5.3 Statistical significance
  • 5.4 A test for differences of sample means: 95% Confidence Intervals
  • 5.5 Error bars in figures
  • 5.6 Discussing statistics in your scientific writing
  • 6 Scatter plot, trendline, and linear regression
  • 7 The t-test of Means
  • 8 Paired t-test
  • 9 Two-Tailed and One-Tailed Tests
  • 10 Variation on t-tests: ANOVA
  • 11 Reporting the Results of a Statistical Test
  • 12 Summary of statistical tests
  • 1 Objectives
  • 2 Project timeline
  • 3 Background
  • 4 Previous work in the BSCI 111 class
  • 5 General notes about the project
  • 6 About the paper
  • 7 References

Nearly all journal articles are divided into the following major sections: abstract, introduction, methods, results, discussion, and references.  Usually the sections are labeled as such, although often the introduction (and sometimes the abstract) is not labeled.  Sometimes alternative section titles are used.  The abstract is sometimes called the "summary", the methods are sometimes called "materials and methods", and the discussion is sometimes called "conclusions".   Some journals also include the minor sections of "key words" following the abstract, and "acknowledgments" following the discussion.  In some journals, the sections may be divided into subsections that are given descriptive titles.  However, the general division into the six major sections is nearly universal.

3.2.1 Abstract

The abstract is a short summary (150-200 words or less) of the important points of the paper.  It does not generally include background information.  There may be a very brief statement of the rationale for conducting the study.  It describes what was done, but without details.  It also describes the results in a summarized way that usually includes whether or not the statistical tests were significant.  It usually concludes with a brief statement of the importance of the results.  Abstracts do not include references.  When writing a paper, the abstract is always the last part to be written.

The purpose of the abstract is to allow potential readers of a paper to find out the important points of the paper without having to actually read the paper.  It should be a self-contained unit capable of being understood without the benefit of the text of the article . It essentially serves as an "advertisement" for the paper that readers use to determine whether or not they actually want to wade through the entire paper or not.  Abstracts are generally freely available in electronic form and are often presented in the results of an electronic search.  If searchers do not have electronic access to the journal in which the article is published, the abstract is the only means that they have to decide whether to go through the effort (going to the library to look up the paper journal, requesting a reprint from the author, buying a copy of the article from a service, requesting the article by Interlibrary Loan) of acquiring the article.  Therefore it is important that the abstract accurately and succinctly presents the most important information in the article.

3.2.2 Introduction

The introduction provides the background information necessary to understand why the described experiment was conducted.  The introduction should describe previous research on the topic that has led to the unanswered questions being addressed by the experiment and should cite important previous papers that form the background for the experiment.  The introduction should also state in an organized fashion the goals of the research, i.e. the particular, specific questions that will be tested in the experiments.  There should be a one-to-one correspondence between questions raised in the introduction and points discussed in the conclusion section of the paper.  In other words, do not raise questions in the introduction unless you are going to have some kind of answer to the question that you intend to discuss at the end of the paper. 

You may have been told that every paper must have a hypothesis that can be clearly stated.  That is often true, but not always.  If your experiment involves a manipulation which tests a specific hypothesis, then you should clearly state that hypothesis.  On the other hand, if your experiment was primarily exploratory, descriptive, or measurative, then you probably did not have an a priori hypothesis, so don't pretend that you did and make one up.  (See the discussion in the introduction to Experiment 4 for more on this.)  If you state a hypothesis in the introduction, it should be a general hypothesis and not a null or alternative hypothesis for a statistical test.  If it is necessary to explain how a statistical test will help you evaluate your general hypothesis, explain that in the methods section. 

A good introduction should be fairly heavy with citations.  This indicates to the reader that the authors are informed about previous work on the topic and are not working in a vacuum.  Citations also provide jumping-off points to allow the reader to explore other tangents to the subject that are not directly addressed in the paper.  If the paper supports or refutes previous work, readers can look up the citations and make a comparison for themselves. 

"Do not get lost in reviewing background information. Remember that the Introduction is meant to introduce the reader to your research, not summarize and evaluate all past literature on the subject (which is the purpose of a review paper). Many of the other studies you may be tempted to discuss in your Introduction are better saved for the Discussion, where they become a powerful tool for comparing and interpreting your results. Include only enough background information to allow your reader to understand why you are asking the questions you are and why your hyptheses are reasonable ones. Often, a brief explanation of the theory involved is sufficient. …

Write this section in the past or present tense, never in the future. " (Steingraber et al. 1985)

3.2.3 Methods (taken verbatim from Steingraber et al. 1985)

The function of this section is to describe all experimental procedures, including controls. The description should be complete enough to enable someone else to repeat your work. If there is more than one part to the experiment, it is a good idea to describe your methods and present your results in the same order in each section. This may not be the same order in which the experiments were performed -it is up to you to decide what order of presentation will make the most sense to your reader.

1. Explain why each procedure was done, i.e., what variable were you measuring and why? Example:

Difficult to understand : First, I removed the frog muscle and then I poured Ringer’s solution on it. Next, I attached it to the kymograph.

Improved: I removed the frog muscle and poured Ringer’s solution on it to prevent it from drying out. I then attached the muscle to the kymograph in order to determine the minimum voltage required for contraction.

2. Experimental procedures and results are narrated in the past tense (what you did, what you found, etc.) whereas conclusions from your results are given in the present tense.

3. Mathematical equations and statistical tests are considered mathematical methods and should be described in this section along with the actual experimental work.

4. Use active rather than passive voice when possible.  [Note: see Section 3.1.4 for more about this.]  Always use the singular "I" rather than the plural "we" when you are the only author of the paper.  Throughout the paper, avoid contractions, e.g. did not vs. didn’t.

5. If any of your methods is fully described in a previous publication (yours or someone else’s), you can cite that instead of describing the procedure again.

Example: The chromosomes were counted at meiosis in the anthers with the standard acetocarmine technique of Snow (1955).

3.2.4 Results (with excerpts from Steingraber et al. 1985)

The function of this section is to summarize general trends in the data without comment, bias, or interpretation. The results of statistical tests applied to your data are reported in this section although conclusions about your original hypotheses are saved for the Discussion section.

Tables and figures should be used when they are a more efficient way to convey information than verbal description. They must be independent units, accompanied by explanatory captions that allow them to be understood by someone who has not read the text. Do not repeat in the text the information in tables and figures, but do cite them, with a summary statement when that is appropriate.  Example:

Incorrect: The results are given in Figure 1.

Correct: Temperature was directly proportional to metabolic rate (Fig. 1).

Please note that the entire word "Figure" is almost never written in an article.  It is nearly always abbreviated as "Fig." and capitalized.  Tables are cited in the same way, although Table is not abbreviated.

Whenever possible, use a figure instead of a table. Relationships between numbers are more readily grasped when they are presented graphically rather than as columns in a table.

Data may be presented in figures and tables, but this may not substitute for a verbal summary of the findings. The text should be understandable by someone who has not seen your figures and tables.

1. All results should be presented, including those that do not support the hypothesis.

2. Statements made in the text must be supported by the results contained in figures and tables.

3. The results of statistical tests can be presented in parentheses following a verbal description.

Example: Fruit size was significantly greater in trees growing alone (t = 3.65, df = 2, p < 0.05).

Simple results of statistical tests may be reported in the text as shown in the preceding example.  The results of multiple tests may be reported in a table if that increases clarity. (See Section 11 of the Statistics Manual for more details about reporting the results of statistical tests.)  It is not necessary to provide a citation for a simple t-test of means, paired t-test, or linear regression.  If you use other tests, you should cite the text or reference you followed to do the test.  In your materials and methods section, you should report how you did the test (e.g. using the statistical analysis package of Excel). 

It is NEVER appropriate to simply paste the results from statistical software into the results section of your paper.  The output generally reports more information than is required and it is not in an appropriate format for a paper.

3.2.4.1 Tables

  • Do not repeat information in a table that you are depicting in a graph or histogram; include a table only if it presents new information.
  • It is easier to compare numbers by reading down a column rather than across a row. Therefore, list sets of data you want your reader to compare in vertical form.
  • Provide each table with a number (Table 1, Table 2, etc.) and a title. The numbered title is placed above the table .
  • Please see Section 11 of the Excel Reference and Statistics Manual for further information on reporting the results of statistical tests.

3.2.4.2. Figures

  • These comprise graphs, histograms, and illustrations, both drawings and photographs. Provide each figure with a number (Fig. 1, Fig. 2, etc.) and a caption (or "legend") that explains what the figure shows. The numbered caption is placed below the figure .  Figure legend = Figure caption.
  • Figures submitted for publication must be "photo ready," i.e., they will appear just as you submit them, or photographically reduced. Therefore, when you graduate from student papers to publishable manuscripts, you must learn to prepare figures that will not embarrass you. At the present time, virtually all journals require manuscripts to be submitted electronically and it is generally assumed that all graphs and maps will be created using software rather than being created by hand.  Nearly all journals have specific guidelines for the file types, resolution, and physical widths required for figures.  Only in a few cases (e.g. sketched diagrams) would figures still be created by hand using ink and those figures would be scanned and labeled using graphics software.  Proportions must be the same as those of the page in the journal to which the paper will be submitted. 
  • Graphs and Histograms: Both can be used to compare two variables. However, graphs show continuous change, whereas histograms show discrete variables only.  You can compare groups of data by plotting two or even three lines on one graph, but avoid cluttered graphs that are hard to read, and do not plot unrelated trends on the same graph. For both graphs, and histograms, plot the independent variable on the horizontal (x) axis and the dependent variable on the vertical (y) axis. Label both axes, including units of measurement except in the few cases where variables are unitless, such as absorbance.
  • Drawings and Photographs: These are used to illustrate organisms, experimental apparatus, models of structures, cellular and subcellular structure, and results of procedures like electrophoresis. Preparing such figures well is a lot of work and can be very expensive, so each figure must add enough to justify its preparation and publication, but good figures can greatly enhance a professional article, as your reading in biological journals has already shown.

3.2.5 Discussion (taken from Steingraber et al. 1985)

The function of this section is to analyze the data and relate them to other studies. To "analyze" means to evaluate the meaning of your results in terms of the original question or hypothesis and point out their biological significance.

1. The Discussion should contain at least:

  • the relationship between the results and the original hypothesis, i.e., whether they support the hypothesis, or cause it to be rejected or modified
  • an integration of your results with those of previous studies in order to arrive at explanations for the observed phenomena
  • possible explanations for unexpected results and observations, phrased as hypotheses that can be tested by realistic experimental procedures, which you should describe

2. Trends that are not statistically significant can still be discussed if they are suggestive or interesting, but cannot be made the basis for conclusions as if they were significant.

3. Avoid redundancy between the Results and the Discussion section. Do not repeat detailed descriptions of the data and results in the Discussion. In some journals, Results and Discussions are joined in a single section, in order to permit a single integrated treatment with minimal repetition. This is more appropriate for short, simple articles than for longer, more complicated ones.

4. End the Discussion with a summary of the principal points you want the reader to remember. This is also the appropriate place to propose specific further study if that will serve some purpose, but do not end with the tired cliché that "this problem needs more study." All problems in biology need more study. Do not close on what you wish you had done, rather finish stating your conclusions and contributions.

3.2.6 Title

The title of the paper should be the last thing that you write.  That is because it should distill the essence of the paper even more than the abstract (the next to last thing that you write). 

The title should contain three elements:

1. the name of the organism studied;

2. the particular aspect or system studied;

3. the variable(s) manipulated.

Do not be afraid to be grammatically creative. Here are some variations on a theme, all suitable as titles:

THE EFFECT OF TEMPERATURE ON GERMINATION OF ZEA MAYS

DOES TEMPERATURE AFFECT GERMINATION OF ZEA MAYS?

TEMPERATURE AND ZEA MAYS GERMINATION: IMPLICATIONS FOR AGRICULTURE

Sometimes it is possible to include the principal result or conclusion in the title:

HIGH TEMPERATURES REDUCE GERMINATION OF ZEA MAYS

Note for the BSCI 1510L class: to make your paper look more like a real paper, you can list all of the other group members as co-authors.  However, if you do that, you should list you name first so that we know that you wrote it.

3.2.7 Literature Cited

Please refer to section 2.1 of this guide.

  • << Previous: 3.1 Specific details regarding scientific writing
  • Next: 4 For further information >>
  • Last Updated: Apr 19, 2023 2:37 PM
  • URL: https://researchguides.library.vanderbilt.edu/bsci1510L

Creative Commons License

Banner

How to Write a Research Paper: Parts of the Paper

  • Choosing Your Topic
  • Citation & Style Guides This link opens in a new window
  • Critical Thinking
  • Evaluating Information
  • Parts of the Paper
  • Writing Tips from UNC-Chapel Hill
  • Librarian Contact

Parts of the Research Paper Papers should have a beginning, a middle, and an end. Your introductory paragraph should grab the reader's attention, state your main idea, and indicate how you will support it. The body of the paper should expand on what you have stated in the introduction. Finally, the conclusion restates the paper's thesis and should explain what you have learned, giving a wrap up of your main ideas.

1. The Title The title should be specific and indicate the theme of the research and what ideas it addresses. Use keywords that help explain your paper's topic to the reader. Try to avoid abbreviations and jargon. Think about keywords that people would use to search for your paper and include them in your title.

2. The Abstract The abstract is used by readers to get a quick overview of your paper. Typically, they are about 200 words in length (120 words minimum to  250 words maximum). The abstract should introduce the topic and thesis, and should provide a general statement about what you have found in your research. The abstract allows you to mention each major aspect of your topic and helps readers decide whether they want to read the rest of the paper. Because it is a summary of the entire research paper, it is often written last. 

3. The Introduction The introduction should be designed to attract the reader's attention and explain the focus of the research. You will introduce your overview of the topic,  your main points of information, and why this subject is important. You can introduce the current understanding and background information about the topic. Toward the end of the introduction, you add your thesis statement, and explain how you will provide information to support your research questions. This provides the purpose and focus for the rest of the paper.

4. Thesis Statement Most papers will have a thesis statement or main idea and supporting facts/ideas/arguments. State your main idea (something of interest or something to be proven or argued for or against) as your thesis statement, and then provide your supporting facts and arguments. A thesis statement is a declarative sentence that asserts the position a paper will be taking. It also points toward the paper's development. This statement should be both specific and arguable. Generally, the thesis statement will be placed at the end of the first paragraph of your paper. The remainder of your paper will support this thesis.

Students often learn to write a thesis as a first step in the writing process, but often, after research, a writer's viewpoint may change. Therefore a thesis statement may be one of the final steps in writing. 

Examples of Thesis Statements from Purdue OWL

5. The Literature Review The purpose of the literature review is to describe past important research and how it specifically relates to the research thesis. It should be a synthesis of the previous literature and the new idea being researched. The review should examine the major theories related to the topic to date and their contributors. It should include all relevant findings from credible sources, such as academic books and peer-reviewed journal articles. You will want  to:

  • Explain how the literature helps the researcher understand the topic.
  • Try to show connections and any disparities between the literature.
  • Identify new ways to interpret prior research.
  • Reveal any gaps that exist in the literature.

More about writing a literature review. . .

6. The Discussion ​The purpose of the discussion is to interpret and describe what you have learned from your research. Make the reader understand why your topic is important. The discussion should always demonstrate what you have learned from your readings (and viewings) and how that learning has made the topic evolve, especially from the short description of main points in the introduction.Explain any new understanding or insights you have had after reading your articles and/or books. Paragraphs should use transitioning sentences to develop how one paragraph idea leads to the next. The discussion will always connect to the introduction, your thesis statement, and the literature you reviewed, but it does not simply repeat or rearrange the introduction. You want to: 

  • Demonstrate critical thinking, not just reporting back facts that you gathered.
  • If possible, tell how the topic has evolved over the past and give it's implications for the future.
  • Fully explain your main ideas with supporting information.
  • Explain why your thesis is correct giving arguments to counter points.

7. The Conclusion A concluding paragraph is a brief summary of your main ideas and restates the paper's main thesis, giving the reader the sense that the stated goal of the paper has been accomplished. What have you learned by doing this research that you didn't know before? What conclusions have you drawn? You may also want to suggest further areas of study, improvement of research possibilities, etc. to demonstrate your critical thinking regarding your research.

  • << Previous: Evaluating Information
  • Next: Research >>
  • Last Updated: Feb 13, 2024 8:35 AM
  • URL: https://libguides.ucc.edu/research_paper

The Structure of a Scientific Paper

Cite this chapter.

Book cover

  • David M. Schultz  

6413 Accesses

1 Citations

Organization is essential for a well-written scientific document. The readers must know where to quickly find the information they seek, from the cover page to the reference list. This chapter explains the parts of a typical scientific document, how to structure these parts into a well-organized document, and how to write each part to effectively communicate the science.

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

Access this chapter

Institutional subscriptions

Unable to display preview.  Download preview PDF.

You can also search for this author in PubMed   Google Scholar

Rights and permissions

Reprints and permissions

Copyright information

© 2009 David M. Schultz

About this chapter

Schultz, D.M. (2009). The Structure of a Scientific Paper. In: Eloquent Science. American Meteorological Society, Boston, MA. https://doi.org/10.1007/978-1-935704-03-4_4

Download citation

DOI : https://doi.org/10.1007/978-1-935704-03-4_4

Publisher Name : American Meteorological Society, Boston, MA

Print ISBN : 978-1-878220-91-2

Online ISBN : 978-1-935704-03-4

eBook Packages : Earth and Environmental Science Earth and Environmental Science (R0)

Share this chapter

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Publish with us

Policies and ethics

Societies and partnerships

American Meteorological Society

  • Find a journal
  • Track your research
  • How To Find Articles with Databases
  • How To Evaluate Articles
  • How To Read A Scientific Paper
  • How To Interpret Data
  • How To Write A Lab Report
  • How To Write A Scientific Paper
  • Get More Help
  • Reference: Encyclopedia, Handbooks & Dictionaries
  • Research Tools: Databases, Protocols & Citation Locators
  • E-Journal Lists by Subject
  • Scholarly vs Popular
  • Search Tips
  • Open Resources
  • E-Journal lists by subject
  • Develop a Research Question

Writing a Scientific Paper

Writing a scientific paper is very similar to writing a lab report. The structure of each is primarily the same, but the purpose of each is different. Lab reports are meant to reflect understanding of the material and learn something new, while scientific papers are meant to contribute knowledge to a field of study.  A scientific paper is broken down into eight sections: title, abstract, introduction, methods, results, discussion, conclusion, and references. 

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

Introduction

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

Methods and Materials

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

Other Useful Sources

  • Guidelines for Writing Scientific Papers
  • How to Write a Scientific Article
  • Writing a Scientific Research Article
  • How to Write a Good Scientific Paper
  • << Previous: How To Write A Lab Report
  • Next: Get More Help >>
  • Last Updated: Mar 8, 2024 2:26 PM
  • URL: https://guides.libraries.indiana.edu/STEM

Social media

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

Additional resources

Featured databases.

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

IU Libraries

  • Diversity Resources
  • About IU Libraries
  • Alumni & Friends
  • Departments & Staff
  • Jobs & Libraries HR
  • Intranet (Staff)
  • IUL site admin
  • Free Materials
  • English Language Editing
  • Technical Scientific Editing
  • Scientific Writing Workshops
  • Online courses
  • Meet the Team

Scientific paper structure: get started with this quick and easy guide

  • by kayciebutler
  • November 13, 2020 November 13, 2020

scientific paper structure is an hourglass

Why the scientific paper structure? It mimics the research process!

Science can be daunting enough – the scientific paper structure doesn’t have to be, too!

In fact, the structure of a scientific paper is meant to be anything but daunting, as it is designed to mimic how science actually progresses.

Don’t believe me? Think about this –

–>Research usually starts with a topic ( title ) .

–>Then, you need to study the s tate of the field around that topic, identify key gaps to address, and form a hypotheses ( introduction ).

–>Next, you gather the tools and equipment you need to do that research ( materials ) and perform experiments ( methods ).

–>After that, you report the results of those experiments ( results ) and see how those results affect the field and integrate back into it ( discussion ).

Helpfully, that is also exactly how your scientific paper is structured.

A scientific research paper is typically ordered:

  • Introduction
  • Materials and Methods
  • Discussion/Conclusions

*Note: This page is going to walk you though the scientific paper structure. If you want info on writing each of these sections, please see my comprehensive page on writing your scientific manuscript !

Scientific paper structure: IMRAD and scope

In more technical terms, the scientific paper is usually structured in what we call the IMRAD format, standing for “ I ntroduction, M ethods, R esults A nd D iscussion.”

An IMRAD-based paper includes:

I ntroduction

  •    Why did you do this research?
  •   What was the original hypothesis?
  •   When, where, and how did you do this research?
  •   What materials or subjects were involved?
  •   What did you discover?
  •   Was the tested hypothesis true?

a nd D iscussion

  •   What do your results mean?
  •   How does this fit within the field?
  •   What are the future  prospectives ?

Besides mimicking the research process, the structure of an IMRAD paper is also helpful for the reader in terms of the the scope of the paper and is designed to draw them in and then show them how your work matters.

What is the scope?

The scope indicates how broadly or narrowly the writing is focused. If the writing in a certain section has a broad scope , it is designed to be accessible to a broad audience . If the writing in a section has a narrow scope , it is designed to be the most focused on your specific work – which is directly accessible to a much smaller audience .

To show you what I mean, I made this diagram that shows how the scope of an IMRAD paper changes from beginning to end:

scientific paper structure is an hourglass

Note how the scope of a scientific paper makes an hourglass shape.

This makes sense, as the important results of your paper are the narrowest in scope. Because this scope is so narrow, it is not widely known, so it would not be accessible to a reader unless it was bookended with information that is much broader in scope, or information that is more well known and understood. This is how you teach the reader what they need to know to understand your work and give them the tools to place your work in context.

Therefore, the introduction of our paper is going to start at the very broadest scope, first introducing the reader to our field in general and then to our research more specifically. In this way, we will start at a very broad scope and slowly narrow into the results – which represent the narrowest scope in our paper.

Scientific paper structure: Key parts

1. title and abstract: attract the reader’s attention.

A scientific paper usually starts with two key parts that help attract a reader’s attention to your work: the title and abstract.

These parts are designed to essentially be the advertisement for your paper.

This means they need to be informative enough about the content of the paper to attract the right readers to your paper, and they also need to be written in a way that is interesting enough to attract those who might not otherwise find your paper.

Also note, basically any reader who gets to the paper body will have read your title and then abstract. By making sure your title and abstract are as attractive as possible, you can get more readers interested in also reading the paper body!

Circles indicating relative views of paper - title gets the most followed by abstract and then paper body

The title contains the key words of the paper, and tries to organize them in a way that lets the reader know what kind of study you conducted and roughly what you accomplished in that paper.

For all of my advice on writing your title, go here .

The abstract is also written to draw attention to your papers, so you want to structure it in a similar hourglass shape as the paper body.

The abstract should

  • start with a broad problem that is relatable to the average reader of that journal
  • indicate how your proposed to solve that problem ( hypothesis or research objectives )
  • give a few lines about what you did in the paper, including key methods and results
  • end with a statement about why your work is important and why it deserves to be published .

This is a lot to ask of a normally 250 word abstract!

Don’t worry – I show you exactly how to do this. For all of my advice on writing your abstract, go here.

Or, you can download your free abstract writing guide here .

2. Introduction: Introduce the reader to your work

After the reader has opened your paper, they need to be introduced to not only your work, but why it matters. This is where the introduction comes in!

Most scientists are good at introducing the literature surrounding their field – which is a big part of the introduction – but struggle to convey the importance or necessity of their work.

Part of this is because many people fail to see the importance of introducing the entire field to the reader to show why it is important to do research in that field.

Therefore, the introduction should start with a very wide overview and include a paragraph at the beginning that introduces the entire field to the reader.

Hourglass shows scope of introduction of scientific research paper

Paragraphs of your introduction

Paragraph 1. The first paragraph of the introduction should answer the question – “Why does my research field exist?” 

Importantly – this paragraph should include a very clear statement of a gap that still exists in the world that your field of research seeks to fill .

Paragraphs 2-3 . Next, it is important to introduce to the reader why your research project exists, which involves the traditional review of relevant literature that most scientists are comfortable writing. These next 1-2 paragraphs should answer the question – “Why does the research in this paper exist?”

Importantly – these paragraphs should include a very clear statement of a gap that still exists in the field that your specific research project seeks to fil l .

Paragraph 4 . The last paragraph of the introduction should give the reader an overview of what to expect in this paper. It should include a typical “Here, we did…” sentence as well as a very short summary of key methods or results.

But we aren’t done yet…

This final paragraph should also end on a sentence that answers the question – “Why does this work matter and deserve to be published?”

The most impactful introductions all end with this forward-thinking statement that helps the reader place the product of your work into context. Don’t underestimate this sentence – getting the “why” into your reader’s head from the beginning can do wonders for their ability to grasp the importance of your work.

For all of my advice on how to write your introduction, go here !

3. Materials and Methods: Tell the reader what you did and how you did it

After setting up why your research projected needed to exist and what you hoped to accomplish, it is time to tell the reader what you did and how you did it.

In terms of text, this section on your materials and methods is the narrowest in scope of all of you paper, as it related to your project alone.

Hourglass shows scope of materials and methods of scientific research paper

In this section, you need provide enough detail that your work could be repeated.

Tell your reader:

  • what materials you used and where you bought them
  • what equipment you used
  • what protocols you followed
  • how you did each experiment
  • how you analyzed your results
  • how you calculated statistics

If you want your work to be considered robust, others need to be able to repeat it.

At this point, your paper should convey what another lab would need to know to copy what you did in this work.

4. Results: Show the reader what you saw

The final section of the narrow scope in your paper is your results, where you tell the reader what you saw in your experiments.

Hourglass shows scope of results of scientific research paper

These paragraphs tell the story of your paper, and should be designed as such.

For the best readability of this section, the results should be structured such that each paragraph :

  • represents one experiment or group of related experiments
  • begins with a topic sentence that tells the reader what you did in that paragraph and why
  • end with a summary statement (1/2 – 1 sentence) telling the reader the main take-home point of that paragraph

The results section should not :

  • Provide extra introductory info only when it is needed to understand the following work and does not apply to the entire paper
  • Provide only enough here such that the reader understands what experiments were done and what the controls were .
  • The reader should not be able to reproduce your experiments from the details in this section
  • Provide only enough for the reader to understand the rest of the paper plus the paragraph-ending summary statement .

For all of my advice on how to write your results, go here !

5. Discussion: Walk the reader through what your results mean and how they affect the field

At the end of the paper, the reader needs to know what your results mean and how they integrate in the field – it is the only way to understand the importance and impact of your work!

For this, the discussion is the opposite of the introduction – it funnels the reader OUT of your work, building on your results to connect your work to the field and society as a whole.

Hourglass shows scope of discussion of scientific research paper

Paragraphs of your discussion

Paragraph 1. The first paragraph briefly summarizes the main results of the paper and directly shows how they address the gap in the field  that was mentioned in the introduction.

Paragraphs 2-4. These middle paragraphs discuss your results. For each paragraph, take one key result and:

  • analyze it – what does it mean?
  • relate it to the field – how does it tie into other work in the field?
  • relate it to the gap – how does it help fill the gap that you discussed in the introduction?
  • speculate beyond the current limits of the field – what new research questions do these results bring up?
  • f uture directions – how can this research be expanded on in the future?

Final paragraph – the conclusion.   The conclusion should never be a summary of the paper – this misses a great opportunity to highlight the importance and impact of your work, and leave the reader with a forward-thinking outlook.

Hourglass shows broad scope of conclusion of scientific research paper

The conclusion does a disservice to your paper if it doesn’t highlight why your work deserves to be published. Make sure it answers:

  • Why should scientists be excited about this work?
  • Why should non-scientists be excited?

For all of my advice on how to write your discussion, go here !

Scientific paper structure: Putting it all together and writing

Now after seeing how a scientific paper is structured and why, you might still be struggling to write the paper…don’t worry, this is completely normal!

Just because we know the structure we need to strive for, it still isn’t easy to translate our work into a paper. This is because the way a paper is structured is designed to help the reader through the process, but it is not necessarily the easiest ordering for writing a paper.

To now learn how to WRITE your scientific paper, you can find all of my advice on that topic here .

Leave a Reply Cancel reply

Your email address will not be published. Required fields are marked *

FREE ABSTRACT E-COURSE

Videos included in scientific abstract email course

Including: -> Detailed breakdowns of ideal abstracts -> Most common mistakes and how to avoid them ->How to WRITE your abstract from scratch ->And all of our best tips , info , and everything you need to know

But we don’t stop there! Joining our community includes: ->Members-only discounts on all of our courses ->Tips for writing , editing , and publishing your science -> First access to all our material direct to your inbox  – NO SPAM !

Here - have a cookie!

Privacy overview.

  • Privacy Policy

Buy Me a Coffee

Research Method

Home » Research Paper – Structure, Examples and Writing Guide

Research Paper – Structure, Examples and Writing Guide

Table of Contents

Research Paper

Research Paper

Definition:

Research Paper is a written document that presents the author’s original research, analysis, and interpretation of a specific topic or issue.

It is typically based on Empirical Evidence, and may involve qualitative or quantitative research methods, or a combination of both. The purpose of a research paper is to contribute new knowledge or insights to a particular field of study, and to demonstrate the author’s understanding of the existing literature and theories related to the topic.

Structure of Research Paper

The structure of a research paper typically follows a standard format, consisting of several sections that convey specific information about the research study. The following is a detailed explanation of the structure of a research paper:

The title page contains the title of the paper, the name(s) of the author(s), and the affiliation(s) of the author(s). It also includes the date of submission and possibly, the name of the journal or conference where the paper is to be published.

The abstract is a brief summary of the research paper, typically ranging from 100 to 250 words. It should include the research question, the methods used, the key findings, and the implications of the results. The abstract should be written in a concise and clear manner to allow readers to quickly grasp the essence of the research.

Introduction

The introduction section of a research paper provides background information about the research problem, the research question, and the research objectives. It also outlines the significance of the research, the research gap that it aims to fill, and the approach taken to address the research question. Finally, the introduction section ends with a clear statement of the research hypothesis or research question.

Literature Review

The literature review section of a research paper provides an overview of the existing literature on the topic of study. It includes a critical analysis and synthesis of the literature, highlighting the key concepts, themes, and debates. The literature review should also demonstrate the research gap and how the current study seeks to address it.

The methods section of a research paper describes the research design, the sample selection, the data collection and analysis procedures, and the statistical methods used to analyze the data. This section should provide sufficient detail for other researchers to replicate the study.

The results section presents the findings of the research, using tables, graphs, and figures to illustrate the data. The findings should be presented in a clear and concise manner, with reference to the research question and hypothesis.

The discussion section of a research paper interprets the findings and discusses their implications for the research question, the literature review, and the field of study. It should also address the limitations of the study and suggest future research directions.

The conclusion section summarizes the main findings of the study, restates the research question and hypothesis, and provides a final reflection on the significance of the research.

The references section provides a list of all the sources cited in the paper, following a specific citation style such as APA, MLA or Chicago.

How to Write Research Paper

You can write Research Paper by the following guide:

  • Choose a Topic: The first step is to select a topic that interests you and is relevant to your field of study. Brainstorm ideas and narrow down to a research question that is specific and researchable.
  • Conduct a Literature Review: The literature review helps you identify the gap in the existing research and provides a basis for your research question. It also helps you to develop a theoretical framework and research hypothesis.
  • Develop a Thesis Statement : The thesis statement is the main argument of your research paper. It should be clear, concise and specific to your research question.
  • Plan your Research: Develop a research plan that outlines the methods, data sources, and data analysis procedures. This will help you to collect and analyze data effectively.
  • Collect and Analyze Data: Collect data using various methods such as surveys, interviews, observations, or experiments. Analyze data using statistical tools or other qualitative methods.
  • Organize your Paper : Organize your paper into sections such as Introduction, Literature Review, Methods, Results, Discussion, and Conclusion. Ensure that each section is coherent and follows a logical flow.
  • Write your Paper : Start by writing the introduction, followed by the literature review, methods, results, discussion, and conclusion. Ensure that your writing is clear, concise, and follows the required formatting and citation styles.
  • Edit and Proofread your Paper: Review your paper for grammar and spelling errors, and ensure that it is well-structured and easy to read. Ask someone else to review your paper to get feedback and suggestions for improvement.
  • Cite your Sources: Ensure that you properly cite all sources used in your research paper. This is essential for giving credit to the original authors and avoiding plagiarism.

Research Paper Example

Note : The below example research paper is for illustrative purposes only and is not an actual research paper. Actual research papers may have different structures, contents, and formats depending on the field of study, research question, data collection and analysis methods, and other factors. Students should always consult with their professors or supervisors for specific guidelines and expectations for their research papers.

Research Paper Example sample for Students:

Title: The Impact of Social Media on Mental Health among Young Adults

Abstract: This study aims to investigate the impact of social media use on the mental health of young adults. A literature review was conducted to examine the existing research on the topic. A survey was then administered to 200 university students to collect data on their social media use, mental health status, and perceived impact of social media on their mental health. The results showed that social media use is positively associated with depression, anxiety, and stress. The study also found that social comparison, cyberbullying, and FOMO (Fear of Missing Out) are significant predictors of mental health problems among young adults.

Introduction: Social media has become an integral part of modern life, particularly among young adults. While social media has many benefits, including increased communication and social connectivity, it has also been associated with negative outcomes, such as addiction, cyberbullying, and mental health problems. This study aims to investigate the impact of social media use on the mental health of young adults.

Literature Review: The literature review highlights the existing research on the impact of social media use on mental health. The review shows that social media use is associated with depression, anxiety, stress, and other mental health problems. The review also identifies the factors that contribute to the negative impact of social media, including social comparison, cyberbullying, and FOMO.

Methods : A survey was administered to 200 university students to collect data on their social media use, mental health status, and perceived impact of social media on their mental health. The survey included questions on social media use, mental health status (measured using the DASS-21), and perceived impact of social media on their mental health. Data were analyzed using descriptive statistics and regression analysis.

Results : The results showed that social media use is positively associated with depression, anxiety, and stress. The study also found that social comparison, cyberbullying, and FOMO are significant predictors of mental health problems among young adults.

Discussion : The study’s findings suggest that social media use has a negative impact on the mental health of young adults. The study highlights the need for interventions that address the factors contributing to the negative impact of social media, such as social comparison, cyberbullying, and FOMO.

Conclusion : In conclusion, social media use has a significant impact on the mental health of young adults. The study’s findings underscore the need for interventions that promote healthy social media use and address the negative outcomes associated with social media use. Future research can explore the effectiveness of interventions aimed at reducing the negative impact of social media on mental health. Additionally, longitudinal studies can investigate the long-term effects of social media use on mental health.

Limitations : The study has some limitations, including the use of self-report measures and a cross-sectional design. The use of self-report measures may result in biased responses, and a cross-sectional design limits the ability to establish causality.

Implications: The study’s findings have implications for mental health professionals, educators, and policymakers. Mental health professionals can use the findings to develop interventions that address the negative impact of social media use on mental health. Educators can incorporate social media literacy into their curriculum to promote healthy social media use among young adults. Policymakers can use the findings to develop policies that protect young adults from the negative outcomes associated with social media use.

References :

  • Twenge, J. M., & Campbell, W. K. (2019). Associations between screen time and lower psychological well-being among children and adolescents: Evidence from a population-based study. Preventive medicine reports, 15, 100918.
  • Primack, B. A., Shensa, A., Escobar-Viera, C. G., Barrett, E. L., Sidani, J. E., Colditz, J. B., … & James, A. E. (2017). Use of multiple social media platforms and symptoms of depression and anxiety: A nationally-representative study among US young adults. Computers in Human Behavior, 69, 1-9.
  • Van der Meer, T. G., & Verhoeven, J. W. (2017). Social media and its impact on academic performance of students. Journal of Information Technology Education: Research, 16, 383-398.

Appendix : The survey used in this study is provided below.

Social Media and Mental Health Survey

  • How often do you use social media per day?
  • Less than 30 minutes
  • 30 minutes to 1 hour
  • 1 to 2 hours
  • 2 to 4 hours
  • More than 4 hours
  • Which social media platforms do you use?
  • Others (Please specify)
  • How often do you experience the following on social media?
  • Social comparison (comparing yourself to others)
  • Cyberbullying
  • Fear of Missing Out (FOMO)
  • Have you ever experienced any of the following mental health problems in the past month?
  • Do you think social media use has a positive or negative impact on your mental health?
  • Very positive
  • Somewhat positive
  • Somewhat negative
  • Very negative
  • In your opinion, which factors contribute to the negative impact of social media on mental health?
  • Social comparison
  • In your opinion, what interventions could be effective in reducing the negative impact of social media on mental health?
  • Education on healthy social media use
  • Counseling for mental health problems caused by social media
  • Social media detox programs
  • Regulation of social media use

Thank you for your participation!

Applications of Research Paper

Research papers have several applications in various fields, including:

  • Advancing knowledge: Research papers contribute to the advancement of knowledge by generating new insights, theories, and findings that can inform future research and practice. They help to answer important questions, clarify existing knowledge, and identify areas that require further investigation.
  • Informing policy: Research papers can inform policy decisions by providing evidence-based recommendations for policymakers. They can help to identify gaps in current policies, evaluate the effectiveness of interventions, and inform the development of new policies and regulations.
  • Improving practice: Research papers can improve practice by providing evidence-based guidance for professionals in various fields, including medicine, education, business, and psychology. They can inform the development of best practices, guidelines, and standards of care that can improve outcomes for individuals and organizations.
  • Educating students : Research papers are often used as teaching tools in universities and colleges to educate students about research methods, data analysis, and academic writing. They help students to develop critical thinking skills, research skills, and communication skills that are essential for success in many careers.
  • Fostering collaboration: Research papers can foster collaboration among researchers, practitioners, and policymakers by providing a platform for sharing knowledge and ideas. They can facilitate interdisciplinary collaborations and partnerships that can lead to innovative solutions to complex problems.

When to Write Research Paper

Research papers are typically written when a person has completed a research project or when they have conducted a study and have obtained data or findings that they want to share with the academic or professional community. Research papers are usually written in academic settings, such as universities, but they can also be written in professional settings, such as research organizations, government agencies, or private companies.

Here are some common situations where a person might need to write a research paper:

  • For academic purposes: Students in universities and colleges are often required to write research papers as part of their coursework, particularly in the social sciences, natural sciences, and humanities. Writing research papers helps students to develop research skills, critical thinking skills, and academic writing skills.
  • For publication: Researchers often write research papers to publish their findings in academic journals or to present their work at academic conferences. Publishing research papers is an important way to disseminate research findings to the academic community and to establish oneself as an expert in a particular field.
  • To inform policy or practice : Researchers may write research papers to inform policy decisions or to improve practice in various fields. Research findings can be used to inform the development of policies, guidelines, and best practices that can improve outcomes for individuals and organizations.
  • To share new insights or ideas: Researchers may write research papers to share new insights or ideas with the academic or professional community. They may present new theories, propose new research methods, or challenge existing paradigms in their field.

Purpose of Research Paper

The purpose of a research paper is to present the results of a study or investigation in a clear, concise, and structured manner. Research papers are written to communicate new knowledge, ideas, or findings to a specific audience, such as researchers, scholars, practitioners, or policymakers. The primary purposes of a research paper are:

  • To contribute to the body of knowledge : Research papers aim to add new knowledge or insights to a particular field or discipline. They do this by reporting the results of empirical studies, reviewing and synthesizing existing literature, proposing new theories, or providing new perspectives on a topic.
  • To inform or persuade: Research papers are written to inform or persuade the reader about a particular issue, topic, or phenomenon. They present evidence and arguments to support their claims and seek to persuade the reader of the validity of their findings or recommendations.
  • To advance the field: Research papers seek to advance the field or discipline by identifying gaps in knowledge, proposing new research questions or approaches, or challenging existing assumptions or paradigms. They aim to contribute to ongoing debates and discussions within a field and to stimulate further research and inquiry.
  • To demonstrate research skills: Research papers demonstrate the author’s research skills, including their ability to design and conduct a study, collect and analyze data, and interpret and communicate findings. They also demonstrate the author’s ability to critically evaluate existing literature, synthesize information from multiple sources, and write in a clear and structured manner.

Characteristics of Research Paper

Research papers have several characteristics that distinguish them from other forms of academic or professional writing. Here are some common characteristics of research papers:

  • Evidence-based: Research papers are based on empirical evidence, which is collected through rigorous research methods such as experiments, surveys, observations, or interviews. They rely on objective data and facts to support their claims and conclusions.
  • Structured and organized: Research papers have a clear and logical structure, with sections such as introduction, literature review, methods, results, discussion, and conclusion. They are organized in a way that helps the reader to follow the argument and understand the findings.
  • Formal and objective: Research papers are written in a formal and objective tone, with an emphasis on clarity, precision, and accuracy. They avoid subjective language or personal opinions and instead rely on objective data and analysis to support their arguments.
  • Citations and references: Research papers include citations and references to acknowledge the sources of information and ideas used in the paper. They use a specific citation style, such as APA, MLA, or Chicago, to ensure consistency and accuracy.
  • Peer-reviewed: Research papers are often peer-reviewed, which means they are evaluated by other experts in the field before they are published. Peer-review ensures that the research is of high quality, meets ethical standards, and contributes to the advancement of knowledge in the field.
  • Objective and unbiased: Research papers strive to be objective and unbiased in their presentation of the findings. They avoid personal biases or preconceptions and instead rely on the data and analysis to draw conclusions.

Advantages of Research Paper

Research papers have many advantages, both for the individual researcher and for the broader academic and professional community. Here are some advantages of research papers:

  • Contribution to knowledge: Research papers contribute to the body of knowledge in a particular field or discipline. They add new information, insights, and perspectives to existing literature and help advance the understanding of a particular phenomenon or issue.
  • Opportunity for intellectual growth: Research papers provide an opportunity for intellectual growth for the researcher. They require critical thinking, problem-solving, and creativity, which can help develop the researcher’s skills and knowledge.
  • Career advancement: Research papers can help advance the researcher’s career by demonstrating their expertise and contributions to the field. They can also lead to new research opportunities, collaborations, and funding.
  • Academic recognition: Research papers can lead to academic recognition in the form of awards, grants, or invitations to speak at conferences or events. They can also contribute to the researcher’s reputation and standing in the field.
  • Impact on policy and practice: Research papers can have a significant impact on policy and practice. They can inform policy decisions, guide practice, and lead to changes in laws, regulations, or procedures.
  • Advancement of society: Research papers can contribute to the advancement of society by addressing important issues, identifying solutions to problems, and promoting social justice and equality.

Limitations of Research Paper

Research papers also have some limitations that should be considered when interpreting their findings or implications. Here are some common limitations of research papers:

  • Limited generalizability: Research findings may not be generalizable to other populations, settings, or contexts. Studies often use specific samples or conditions that may not reflect the broader population or real-world situations.
  • Potential for bias : Research papers may be biased due to factors such as sample selection, measurement errors, or researcher biases. It is important to evaluate the quality of the research design and methods used to ensure that the findings are valid and reliable.
  • Ethical concerns: Research papers may raise ethical concerns, such as the use of vulnerable populations or invasive procedures. Researchers must adhere to ethical guidelines and obtain informed consent from participants to ensure that the research is conducted in a responsible and respectful manner.
  • Limitations of methodology: Research papers may be limited by the methodology used to collect and analyze data. For example, certain research methods may not capture the complexity or nuance of a particular phenomenon, or may not be appropriate for certain research questions.
  • Publication bias: Research papers may be subject to publication bias, where positive or significant findings are more likely to be published than negative or non-significant findings. This can skew the overall findings of a particular area of research.
  • Time and resource constraints: Research papers may be limited by time and resource constraints, which can affect the quality and scope of the research. Researchers may not have access to certain data or resources, or may be unable to conduct long-term studies due to practical limitations.

About the author

' src=

Muhammad Hassan

Researcher, Academic Writer, Web developer

You may also like

Research Paper Citation

How to Cite Research Paper – All Formats and...

Data collection

Data Collection – Methods Types and Examples

Delimitations

Delimitations in Research – Types, Examples and...

Research Paper Formats

Research Paper Format – Types, Examples and...

Research Process

Research Process – Steps, Examples and Tips

Research Design

Research Design – Types, Methods and Examples

Have a language expert improve your writing

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

  • Knowledge Base
  • Research paper

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

Published on September 24, 2022 by Jack Caulfield . Revised on March 27, 2023.

Writing a Research Paper Introduction

The introduction to a research paper is where you set up your topic and approach for the reader. It has several key goals:

  • Present your topic and get the reader interested
  • Provide background or summarize existing research
  • Position your own approach
  • Detail your specific research problem and problem statement
  • Give an overview of the paper’s structure

The introduction looks slightly different depending on whether your paper presents the results of original empirical research or constructs an argument by engaging with a variety of sources.

Instantly correct all language mistakes in your text

Upload your document to correct all your mistakes in minutes

upload-your-document-ai-proofreader

Table of contents

Step 1: introduce your topic, step 2: describe the background, step 3: establish your research problem, step 4: specify your objective(s), step 5: map out your paper, research paper introduction examples, frequently asked questions about the research paper introduction.

The first job of the introduction is to tell the reader what your topic is and why it’s interesting or important. This is generally accomplished with a strong opening hook.

The hook is a striking opening sentence that clearly conveys the relevance of your topic. Think of an interesting fact or statistic, a strong statement, a question, or a brief anecdote that will get the reader wondering about your topic.

For example, the following could be an effective hook for an argumentative paper about the environmental impact of cattle farming:

A more empirical paper investigating the relationship of Instagram use with body image issues in adolescent girls might use the following hook:

Don’t feel that your hook necessarily has to be deeply impressive or creative. Clarity and relevance are still more important than catchiness. The key thing is to guide the reader into your topic and situate your ideas.

The only proofreading tool specialized in correcting academic writing - try for free!

The academic proofreading tool has been trained on 1000s of academic texts and by native English editors. Making it the most accurate and reliable proofreading tool for students.

parts of a science research paper

Try for free

This part of the introduction differs depending on what approach your paper is taking.

In a more argumentative paper, you’ll explore some general background here. In a more empirical paper, this is the place to review previous research and establish how yours fits in.

Argumentative paper: Background information

After you’ve caught your reader’s attention, specify a bit more, providing context and narrowing down your topic.

Provide only the most relevant background information. The introduction isn’t the place to get too in-depth; if more background is essential to your paper, it can appear in the body .

Empirical paper: Describing previous research

For a paper describing original research, you’ll instead provide an overview of the most relevant research that has already been conducted. This is a sort of miniature literature review —a sketch of the current state of research into your topic, boiled down to a few sentences.

This should be informed by genuine engagement with the literature. Your search can be less extensive than in a full literature review, but a clear sense of the relevant research is crucial to inform your own work.

Begin by establishing the kinds of research that have been done, and end with limitations or gaps in the research that you intend to respond to.

The next step is to clarify how your own research fits in and what problem it addresses.

Argumentative paper: Emphasize importance

In an argumentative research paper, you can simply state the problem you intend to discuss, and what is original or important about your argument.

Empirical paper: Relate to the literature

In an empirical research paper, try to lead into the problem on the basis of your discussion of the literature. Think in terms of these questions:

  • What research gap is your work intended to fill?
  • What limitations in previous work does it address?
  • What contribution to knowledge does it make?

You can make the connection between your problem and the existing research using phrases like the following.

Now you’ll get into the specifics of what you intend to find out or express in your research paper.

The way you frame your research objectives varies. An argumentative paper presents a thesis statement, while an empirical paper generally poses a research question (sometimes with a hypothesis as to the answer).

Argumentative paper: Thesis statement

The thesis statement expresses the position that the rest of the paper will present evidence and arguments for. It can be presented in one or two sentences, and should state your position clearly and directly, without providing specific arguments for it at this point.

Empirical paper: Research question and hypothesis

The research question is the question you want to answer in an empirical research paper.

Present your research question clearly and directly, with a minimum of discussion at this point. The rest of the paper will be taken up with discussing and investigating this question; here you just need to express it.

A research question can be framed either directly or indirectly.

  • This study set out to answer the following question: What effects does daily use of Instagram have on the prevalence of body image issues among adolescent girls?
  • We investigated the effects of daily Instagram use on the prevalence of body image issues among adolescent girls.

If your research involved testing hypotheses , these should be stated along with your research question. They are usually presented in the past tense, since the hypothesis will already have been tested by the time you are writing up your paper.

For example, the following hypothesis might respond to the research question above:

Here's why students love Scribbr's proofreading services

Discover proofreading & editing

The final part of the introduction is often dedicated to a brief overview of the rest of the paper.

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.

  • This paper will first discuss several examples of survey-based research into adolescent social media use, then will go on to …
  • This paper first discusses several examples of survey-based research into adolescent social media use, then goes on to …

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

  • Argumentative paper
  • Empirical paper

Are cows responsible for climate change? A recent study (RIVM, 2019) shows that cattle farmers account for two thirds of agricultural nitrogen emissions in the Netherlands. These emissions result from nitrogen in manure, which can degrade into ammonia and enter the atmosphere. The study’s calculations show that agriculture is the main source of nitrogen pollution, accounting for 46% of the country’s total emissions. By comparison, road traffic and households are responsible for 6.1% each, the industrial sector for 1%. While efforts are being made to mitigate these emissions, policymakers are reluctant to reckon with the scale of the problem. The approach presented here is a radical one, but commensurate with the issue. This paper argues that the Dutch government must stimulate and subsidize livestock farmers, especially cattle farmers, to transition to sustainable vegetable farming. It first establishes the inadequacy of current mitigation measures, then discusses the various advantages of the results proposed, and finally addresses potential objections to the plan on economic grounds.

The rise of social media has been accompanied by a sharp increase in the prevalence of body image issues among women and girls. This correlation has received significant academic attention: Various empirical studies have been conducted into Facebook usage among adolescent girls (Tiggermann & Slater, 2013; Meier & Gray, 2014). These studies have consistently found that the visual and interactive aspects of the platform have the greatest influence on body image issues. Despite this, highly visual social media (HVSM) such as Instagram have yet to be robustly researched. This paper sets out to address this research gap. We investigated the effects of daily Instagram use on the prevalence of body image issues among adolescent girls. It was hypothesized that daily Instagram use would be associated with an increase in body image concerns and a decrease in self-esteem ratings.

The introduction of a research paper includes several key elements:

  • A hook to catch the reader’s interest
  • Relevant background on the topic
  • Details of your research problem

and your problem statement

  • A thesis statement or research question
  • Sometimes an overview of the paper

Don’t feel that you have to write the introduction first. The introduction is often one of the last parts of the research paper you’ll write, along with the conclusion.

This is because it can be easier to introduce your paper once you’ve already written the body ; you may not have the clearest idea of your arguments until you’ve written them, and things can change during the writing process .

The way you present your research problem in your introduction varies depending on the nature of your research paper . A research paper that presents a sustained argument will usually encapsulate this argument in a thesis statement .

A research paper designed to present the results of empirical research tends to present a research question that it seeks to answer. It may also include a hypothesis —a prediction that will be confirmed or disproved by your research.

Cite this Scribbr article

If you want to cite this source, you can copy and paste the citation or click the “Cite this Scribbr article” button to automatically add the citation to our free Citation Generator.

Caulfield, J. (2023, March 27). Writing a Research Paper Introduction | Step-by-Step Guide. Scribbr. Retrieved April 17, 2024, from https://www.scribbr.com/research-paper/research-paper-introduction/

Is this article helpful?

Jack Caulfield

Jack Caulfield

Other students also liked, writing strong research questions | criteria & examples, writing a research paper conclusion | step-by-step guide, research paper format | apa, mla, & chicago templates, "i thought ai proofreading was useless but..".

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

  • Foundations
  • Write Paper

Search form

  • Experiments
  • Anthropology
  • Self-Esteem
  • Social Anxiety

parts of a science research paper

  • Research Paper >

Parts of a Research Paper

One of the most important aspects of science is ensuring that you get all the parts of the written research paper in the right order.

This article is a part of the guide:

  • Outline Examples
  • Example of a Paper
  • Write a Hypothesis
  • Introduction

Browse Full Outline

  • 1 Write a Research Paper
  • 2 Writing a Paper
  • 3.1 Write an Outline
  • 3.2 Outline Examples
  • 4.1 Thesis Statement
  • 4.2 Write a Hypothesis
  • 5.2 Abstract
  • 5.3 Introduction
  • 5.4 Methods
  • 5.5 Results
  • 5.6 Discussion
  • 5.7 Conclusion
  • 5.8 Bibliography
  • 6.1 Table of Contents
  • 6.2 Acknowledgements
  • 6.3 Appendix
  • 7.1 In Text Citations
  • 7.2 Footnotes
  • 7.3.1 Floating Blocks
  • 7.4 Example of a Paper
  • 7.5 Example of a Paper 2
  • 7.6.1 Citations
  • 7.7.1 Writing Style
  • 7.7.2 Citations
  • 8.1.1 Sham Peer Review
  • 8.1.2 Advantages
  • 8.1.3 Disadvantages
  • 8.2 Publication Bias
  • 8.3.1 Journal Rejection
  • 9.1 Article Writing
  • 9.2 Ideas for Topics

You may have finished the best research project on earth but, if you do not write an interesting and well laid out paper, then nobody is going to take your findings seriously.

The main thing to remember with any research paper is that it is based on an hourglass structure. It begins with general information and undertaking a literature review , and becomes more specific as you nail down a research problem and hypothesis .

Finally, it again becomes more general as you try to apply your findings to the world at general.

Whilst there are a few differences between the various disciplines, with some fields placing more emphasis on certain parts than others, there is a basic underlying structure.

These steps are the building blocks of constructing a good research paper. This section outline how to lay out the parts of a research paper, including the various experimental methods and designs.

The principles for literature review and essays of all types follow the same basic principles.

Reference List

parts of a science research paper

For many students, writing the introduction is the first part of the process, setting down the direction of the paper and laying out exactly what the research paper is trying to achieve.

For others, the introduction is the last thing written, acting as a quick summary of the paper. As long as you have planned a good structure for the parts of a research paper, both approaches are acceptable and it is a matter of preference.

A good introduction generally consists of three distinct parts:

  • You should first give a general presentation of the research problem.
  • You should then lay out exactly what you are trying to achieve with this particular research project.
  • You should then state your own position.

Ideally, you should try to give each section its own paragraph, but this will vary given the overall length of the paper.

1) General Presentation

Look at the benefits to be gained by the research or why the problem has not been solved yet. Perhaps nobody has thought about it, or maybe previous research threw up some interesting leads that the previous researchers did not follow up.

Another researcher may have uncovered some interesting trends, but did not manage to reach the significance level , due to experimental error or small sample sizes .

2) Purpose of the Paper

The research problem does not have to be a statement, but must at least imply what you are trying to find.

Many writers prefer to place the thesis statement or hypothesis here, which is perfectly acceptable, but most include it in the last sentences of the introduction, to give the reader a fuller picture.

3) A Statement of Intent From the Writer

The idea is that somebody will be able to gain an overall view of the paper without needing to read the whole thing. Literature reviews are time-consuming enough, so give the reader a concise idea of your intention before they commit to wading through pages of background.

In this section, you look to give a context to the research, including any relevant information learned during your literature review. You are also trying to explain why you chose this area of research, attempting to highlight why it is necessary. The second part should state the purpose of the experiment and should include the research problem. The third part should give the reader a quick summary of the form that the parts of the research paper is going to take and should include a condensed version of the discussion.

parts of a science research paper

This should be the easiest part of the paper to write, as it is a run-down of the exact design and methodology used to perform the research. Obviously, the exact methodology varies depending upon the exact field and type of experiment .

There is a big methodological difference between the apparatus based research of the physical sciences and the methods and observation methods of social sciences. However, the key is to ensure that another researcher would be able to replicate the experiment to match yours as closely as possible, but still keeping the section concise.

You can assume that anybody reading your paper is familiar with the basic methods, so try not to explain every last detail. For example, an organic chemist or biochemist will be familiar with chromatography, so you only need to highlight the type of equipment used rather than explaining the whole process in detail.

In the case of a survey , if you have too many questions to cover in the method, you can always include a copy of the questionnaire in the appendix . In this case, make sure that you refer to it.

This is probably the most variable part of any research paper, and depends on the results and aims of the experiment.

For quantitative research , it is a presentation of the numerical results and data, whereas for qualitative research it should be a broader discussion of trends, without going into too much detail.

For research generating a lot of results , then it is better to include tables or graphs of the analyzed data and leave the raw data in the appendix, so that a researcher can follow up and check your calculations.

A commentary is essential to linking the results together, rather than just displaying isolated and unconnected charts and figures.

It can be quite difficult to find a good balance between the results and the discussion section, because some findings, especially in a quantitative or descriptive experiment , will fall into a grey area. Try to avoid repeating yourself too often.

It is best to try to find a middle path, where you give a general overview of the data and then expand on it in the discussion - you should try to keep your own opinions and interpretations out of the results section, saving that for the discussion later on.

This is where you elaborate on your findings, and explain what you found, adding your own personal interpretations.

Ideally, you should link the discussion back to the introduction, addressing each point individually.

It’s important to make sure that every piece of information in your discussion is directly related to the thesis statement , or you risk cluttering your findings. In keeping with the hourglass principle, you can expand on the topic later in the conclusion .

The conclusion is where you build on your discussion and try to relate your findings to other research and to the world at large.

In a short research paper, it may be a paragraph or two, or even a few lines.

In a dissertation, it may well be the most important part of the entire paper - not only does it describe the results and discussion in detail, it emphasizes the importance of the results in the field, and ties it in with the previous research.

Some research papers require a recommendations section, postulating the further directions of the research, as well as highlighting how any flaws affected the results. In this case, you should suggest any improvements that could be made to the research design .

No paper is complete without a reference list , documenting all the sources that you used for your research. This should be laid out according to APA , MLA or other specified format, allowing any interested researcher to follow up on the research.

One habit that is becoming more common, especially with online papers, is to include a reference to your own paper on the final page. Lay this out in MLA, APA and Chicago format, allowing anybody referencing your paper to copy and paste it.

  • Psychology 101
  • Flags and Countries
  • Capitals and Countries

Martyn Shuttleworth (Jun 5, 2009). Parts of a Research Paper. Retrieved Apr 17, 2024 from Explorable.com: https://explorable.com/parts-of-a-research-paper

You Are Allowed To Copy The Text

The text in this article is licensed under the Creative Commons-License Attribution 4.0 International (CC BY 4.0) .

This means you're free to copy, share and adapt any parts (or all) of the text in the article, as long as you give appropriate credit and provide a link/reference to this page.

That is it. You don't need our permission to copy the article; just include a link/reference back to this page. You can use it freely (with some kind of link), and we're also okay with people reprinting in publications like books, blogs, newsletters, course-material, papers, wikipedia and presentations (with clear attribution).

Want to stay up to date? Follow us!

Check out the official book.

Learn how to construct, style and format an Academic paper and take your skills to the next level.

parts of a science research paper

(also available as ebook )

Save this course for later

Don't have time for it all now? No problem, save it as a course and come back to it later.

Footer bottom

  • Privacy Policy

parts of a science research paper

  • Subscribe to our RSS Feed
  • Like us on Facebook
  • Follow us on Twitter

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
  • PLoS Comput Biol
  • v.16(7); 2020 Jul

Logo of ploscomp

Ten simple rules for reading a scientific paper

Maureen a. carey.

Division of Infectious Diseases and International Health, Department of Medicine, University of Virginia School of Medicine, Charlottesville, Virginia, United States of America

Kevin L. Steiner

William a. petri, jr, introduction.

“There is no problem that a library card can't solve” according to author Eleanor Brown [ 1 ]. This advice is sound, probably for both life and science, but even the best tool (like the library) is most effective when accompanied by instructions and a basic understanding of how and when to use it.

For many budding scientists, the first day in a new lab setting often involves a stack of papers, an email full of links to pertinent articles, or some promise of a richer understanding so long as one reads enough of the scientific literature. However, the purpose and approach to reading a scientific article is unlike that of reading a news story, novel, or even a textbook and can initially seem unapproachable. Having good habits for reading scientific literature is key to setting oneself up for success, identifying new research questions, and filling in the gaps in one’s current understanding; developing these good habits is the first crucial step.

Advice typically centers around two main tips: read actively and read often. However, active reading, or reading with an intent to understand, is both a learned skill and a level of effort. Although there is no one best way to do this, we present 10 simple rules, relevant to novices and seasoned scientists alike, to teach our strategy for active reading based on our experience as readers and as mentors of undergraduate and graduate researchers, medical students, fellows, and early career faculty. Rules 1–5 are big picture recommendations. Rules 6–8 relate to philosophy of reading. Rules 9–10 guide the “now what?” questions one should ask after reading and how to integrate what was learned into one’s own science.

Rule 1: Pick your reading goal

What you want to get out of an article should influence your approach to reading it. Table 1 includes a handful of example intentions and how you might prioritize different parts of the same article differently based on your goals as a reader.

1 Yay! Welcome!

2 A journal club is when a group of scientists get together to discuss a paper. Usually one person leads the discussion and presents all of the data. The group discusses their own interpretations and the authors’ interpretation.

Rule 2: Understand the author’s goal

In written communication, the reader and the writer are equally important. Both influence the final outcome: in this case, your scientific understanding! After identifying your goal, think about the author’s goal for sharing this project. This will help you interpret the data and understand the author’s interpretation of the data. However, this requires some understanding of who the author(s) are (e.g., what are their scientific interests?), the scientific field in which they work (e.g., what techniques are available in this field?), and how this paper fits into the author’s research (e.g., is this work building on an author’s longstanding project or controversial idea?). This information may be hard to glean without experience and a history of reading. But don’t let this be a discouragement to starting the process; it is by the act of reading that this experience is gained!

A good step toward understanding the goal of the author(s) is to ask yourself: What kind of article is this? Journals publish different types of articles, including methods, review, commentary, resources, and research articles as well as other types that are specific to a particular journal or groups of journals. These article types have different formatting requirements and expectations for content. Knowing the article type will help guide your evaluation of the information presented. Is the article a methods paper, presenting a new technique? Is the article a review article, intended to summarize a field or problem? Is it a commentary, intended to take a stand on a controversy or give a big picture perspective on a problem? Is it a resource article, presenting a new tool or data set for others to use? Is it a research article, written to present new data and the authors’ interpretation of those data? The type of paper, and its intended purpose, will get you on your way to understanding the author’s goal.

Rule 3: Ask six questions

When reading, ask yourself: (1) What do the author(s) want to know (motivation)? (2) What did they do (approach/methods)? (3) Why was it done that way (context within the field)? (4) What do the results show (figures and data tables)? (5) How did the author(s) interpret the results (interpretation/discussion)? (6) What should be done next? (Regarding this last question, the author(s) may provide some suggestions in the discussion, but the key is to ask yourself what you think should come next.)

Each of these questions can and should be asked about the complete work as well as each table, figure, or experiment within the paper. Early on, it can take a long time to read one article front to back, and this can be intimidating. Break down your understanding of each section of the work with these questions to make the effort more manageable.

Rule 4: Unpack each figure and table

Scientists write original research papers primarily to present new data that may change or reinforce the collective knowledge of a field. Therefore, the most important parts of this type of scientific paper are the data. Some people like to scrutinize the figures and tables (including legends) before reading any of the “main text”: because all of the important information should be obtained through the data. Others prefer to read through the results section while sequentially examining the figures and tables as they are addressed in the text. There is no correct or incorrect approach: Try both to see what works best for you. The key is making sure that one understands the presented data and how it was obtained.

For each figure, work to understand each x- and y-axes, color scheme, statistical approach (if one was used), and why the particular plotting approach was used. For each table, identify what experimental groups and variables are presented. Identify what is shown and how the data were collected. This is typically summarized in the legend or caption but often requires digging deeper into the methods: Do not be afraid to refer back to the methods section frequently to ensure a full understanding of how the presented data were obtained. Again, ask the questions in Rule 3 for each figure or panel and conclude with articulating the “take home” message.

Rule 5: Understand the formatting intentions

Just like the overall intent of the article (discussed in Rule 2), the intent of each section within a research article can guide your interpretation. Some sections are intended to be written as objective descriptions of the data (i.e., the Results section), whereas other sections are intended to present the author’s interpretation of the data. Remember though that even “objective” sections are written by and, therefore, influenced by the authors interpretations. Check out Table 2 to understand the intent of each section of a research article. When reading a specific paper, you can also refer to the journal’s website to understand the formatting intentions. The “For Authors” section of a website will have some nitty gritty information that is less relevant for the reader (like word counts) but will also summarize what the journal editors expect in each section. This will help to familiarize you with the goal of each article section.

Research articles typically contain each of these sections, although sometimes the “results” and “discussion” sections (or “discussion” and “conclusion” sections) are merged into one section. Additional sections may be included, based on request of the journal or the author(s). Keep in mind: If it was included, someone thought it was important for you to read.

Rule 6: Be critical

Published papers are not truths etched in stone. Published papers in high impact journals are not truths etched in stone. Published papers by bigwigs in the field are not truths etched in stone. Published papers that seem to agree with your own hypothesis or data are not etched in stone. Published papers that seem to refute your hypothesis or data are not etched in stone.

Science is a never-ending work in progress, and it is essential that the reader pushes back against the author’s interpretation to test the strength of their conclusions. Everyone has their own perspective and may interpret the same data in different ways. Mistakes are sometimes published, but more often these apparent errors are due to other factors such as limitations of a methodology and other limits to generalizability (selection bias, unaddressed, or unappreciated confounders). When reading a paper, it is important to consider if these factors are pertinent.

Critical thinking is a tough skill to learn but ultimately boils down to evaluating data while minimizing biases. Ask yourself: Are there other, equally likely, explanations for what is observed? In addition to paying close attention to potential biases of the study or author(s), a reader should also be alert to one’s own preceding perspective (and biases). Take time to ask oneself: Do I find this paper compelling because it affirms something I already think (or wish) is true? Or am I discounting their findings because it differs from what I expect or from my own work?

The phenomenon of a self-fulfilling prophecy, or expectancy, is well studied in the psychology literature [ 2 ] and is why many studies are conducted in a “blinded” manner [ 3 ]. It refers to the idea that a person may assume something to be true and their resultant behavior aligns to make it true. In other words, as humans and scientists, we often find exactly what we are looking for. A scientist may only test their hypotheses and fail to evaluate alternative hypotheses; perhaps, a scientist may not be aware of alternative, less biased ways to test her or his hypothesis that are typically used in different fields. Individuals with different life, academic, and work experiences may think of several alternative hypotheses, all equally supported by the data.

Rule 7: Be kind

The author(s) are human too. So, whenever possible, give them the benefit of the doubt. An author may write a phrase differently than you would, forcing you to reread the sentence to understand it. Someone in your field may neglect to cite your paper because of a reference count limit. A figure panel may be misreferenced as Supplemental Fig 3E when it is obviously Supplemental Fig 4E. While these things may be frustrating, none are an indication that the quality of work is poor. Try to avoid letting these minor things influence your evaluation and interpretation of the work.

Similarly, if you intend to share your critique with others, be extra kind. An author (especially the lead author) may invest years of their time into a single paper. Hearing a kindly phrased critique can be difficult but constructive. Hearing a rude, brusque, or mean-spirited critique can be heartbreaking, especially for young scientists or those seeking to establish their place within a field and who may worry that they do not belong.

Rule 8: Be ready to go the extra mile

To truly understand a scientific work, you often will need to look up a term, dig into the supplemental materials, or read one or more of the cited references. This process takes time. Some advisors recommend reading an article three times: The first time, simply read without the pressure of understanding or critiquing the work. For the second time, aim to understand the paper. For the third read through, take notes.

Some people engage with a paper by printing it out and writing all over it. The reader might write question marks in the margins to mark parts (s)he wants to return to, circle unfamiliar terms (and then actually look them up!), highlight or underline important statements, and draw arrows linking figures and the corresponding interpretation in the discussion. Not everyone needs a paper copy to engage in the reading process but, whatever your version of “printing it out” is, do it.

Rule 9: Talk about it

Talking about an article in a journal club or more informal environment forces active reading and participation with the material. Studies show that teaching is one of the best ways to learn and that teachers learn the material even better as the teaching task becomes more complex [ 4 – 5 ]; anecdotally, such observations inspired the phrase “to teach is to learn twice.”

Beyond formal settings such as journal clubs, lab meetings, and academic classes, discuss papers with your peers, mentors, and colleagues in person or electronically. Twitter and other social media platforms have become excellent resources for discussing papers with other scientists, the public or your nonscientist friends, or even the paper’s author(s). Describing a paper can be done at multiple levels and your description can contain all of the scientific details, only the big picture summary, or perhaps the implications for the average person in your community. All of these descriptions will solidify your understanding, while highlighting gaps in your knowledge and informing those around you.

Rule 10: Build on it

One approach we like to use for communicating how we build on the scientific literature is by starting research presentations with an image depicting a wall of Lego bricks. Each brick is labeled with the reference for a paper, and the wall highlights the body of literature on which the work is built. We describe the work and conclusions of each paper represented by a labeled brick and discuss each brick and the wall as a whole. The top brick on the wall is left blank: We aspire to build on this work and label this brick with our own work. We then delve into our own research, discoveries, and the conclusions it inspires. We finish our presentations with the image of the Legos and summarize our presentation on that empty brick.

Whether you are reading an article to understand a new topic area or to move a research project forward, effective learning requires that you integrate knowledge from multiple sources (“click” those Lego bricks together) and build upwards. Leveraging published work will enable you to build a stronger and taller structure. The first row of bricks is more stable once a second row is assembled on top of it and so on and so forth. Moreover, the Lego construction will become taller and larger if you build upon the work of others, rather than using only your own bricks.

Build on the article you read by thinking about how it connects to ideas described in other papers and within own work, implementing a technique in your own research, or attempting to challenge or support the hypothesis of the author(s) with a more extensive literature review. Integrate the techniques and scientific conclusions learned from an article into your own research or perspective in the classroom or research lab. You may find that this process strengthens your understanding, leads you toward new and unexpected interests or research questions, or returns you back to the original article with new questions and critiques of the work. All of these experiences are part of the “active reading”: process and are signs of a successful reading experience.

In summary, practice these rules to learn how to read a scientific article, keeping in mind that this process will get easier (and faster) with experience. We are firm believers that an hour in the library will save a week at the bench; this diligent practice will ultimately make you both a more knowledgeable and productive scientist. As you develop the skills to read an article, try to also foster good reading and learning habits for yourself (recommendations here: [ 6 ] and [ 7 ], respectively) and in others. Good luck and happy reading!

Acknowledgments

Thank you to the mentors, teachers, and students who have shaped our thoughts on reading, learning, and what science is all about.

Funding Statement

MAC was supported by the PhRMA Foundation's Postdoctoral Fellowship in Translational Medicine and Therapeutics and the University of Virginia's Engineering-in-Medicine seed grant, and KLS was supported by the NIH T32 Global Biothreats Training Program at the University of Virginia (AI055432). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Penn State University Libraries

Computer science and engineering.

  • Reference Sources
  • Finding Articles and Databases
  • Finding Books
  • Finding Websites
  • Penn State Resources and Organizations
  • Books, Articles, and Other Educational Resources
  • Research Tips
  • Main Parts of a Scientific/Technical Paper
  • Technical Writing Resources
  • Ten Tips for Technical Writing
  • Professional Organizations

Parts of a Technical Paper

The basic parts of a scientific or technical paper are:

Title and Author Information Abstract Introduction Literature Review Methods Results Discussion Conclusions References and Appendices

Plan, Prioritize, Pass!

Get ready for finals with the arc.

Academic Resource Center

How to read and understand a scientific paper

How to read and understand a scientific paper: a guide for non-scientists, london school of economics and political science, jennifer raff.

From vaccinations to climate change, getting science wrong has very real consequences. But journal articles, a primary way science is communicated in academia, are a different format to newspaper articles or blogs and require a level of skill and undoubtedly a greater amount of patience. Here  Jennifer Raff   has prepared a helpful guide for non-scientists on how to read a scientific paper. These steps and tips will be useful to anyone interested in the presentation of scientific findings and raise important points for scientists to consider with their own writing practice.

My post,  The truth about vaccinations: Your physician knows more than the University of Google  sparked a very lively discussion, with comments from several people trying to persuade me (and the other readers) that  their  paper disproved everything that I’d been saying. While I encourage you to go read the comments and contribute your own, here I want to focus on the much larger issue that this debate raised: what constitutes scientific authority?

It’s not just a fun academic problem. Getting the science wrong has very real consequences. For example, when a community doesn’t vaccinate children because they’re afraid of “toxins” and think that prayer (or diet, exercise, and “clean living”) is enough to prevent infection, outbreaks happen.

“Be skeptical. But when you get proof, accept proof.” –Michael Specter

What constitutes enough proof? Obviously everyone has a different answer to that question. But to form a truly educated opinion on a scientific subject, you need to become familiar with current research in that field. And to do that, you have to read the “primary research literature” (often just called “the literature”). You might have tried to read scientific papers before and been frustrated by the dense, stilted writing and the unfamiliar jargon. I remember feeling this way!  Reading and understanding research papers is a skill which every single doctor and scientist has had to learn during graduate school.  You can learn it too, but like any skill it takes patience and practice.

I want to help people become more scientifically literate, so I wrote this guide for how a layperson can approach reading and understanding a scientific research paper. It’s appropriate for someone who has no background whatsoever in science or medicine, and based on the assumption that he or she is doing this for the purpose of getting a  basic  understanding of a paper and deciding whether or not it’s a reputable study.

The type of scientific paper I’m discussing here is referred to as a  primary research article . It’s a peer-reviewed report of new research on a specific question (or questions). Another useful type of publication is a  review article . Review articles are also peer-reviewed, and don’t present new information, but summarize multiple primary research articles, to give a sense of the consensus, debates, and unanswered questions within a field.  (I’m not going to say much more about them here, but be cautious about which review articles you read. Remember that they are only a snapshot of the research at the time they are published.  A review article on, say, genome-wide association studies from 2001 is not going to be very informative in 2013. So much research has been done in the intervening years that the field has changed considerably).

Before you begin: some general advice

Reading a scientific paper is a completely different process than reading an article about science in a blog or newspaper. Not only do you read the sections in a different order than they’re presented, but you also have to take notes, read it multiple times, and probably go look up other papers for some of the details. Reading a single paper may take you a very long time at first. Be patient with yourself. The process will go much faster as you gain experience.

Most primary research papers will be divided into the following sections: Abstract, Introduction, Methods, Results, and Conclusions/Interpretations/Discussion. The order will depend on which journal it’s published in. Some journals have additional files (called Supplementary Online Information) which contain important details of the research, but are published online instead of in the article itself (make sure you don’t skip these files).

Before you begin reading, take note of the authors and their institutional affiliations. Some institutions (e.g. University of Texas) are well-respected; others (e.g.  the Discovery Institute ) may appear to be legitimate research institutions but are actually agenda-driven.  Tip:  g oogle  “Discovery Institute” to see why you don’t want to use it as a scientific authority on evolutionary theory.

Also take note of the journal in which it’s published. Reputable (biomedical) journals will be indexed by  Pubmed . [EDIT: Several people have reminded me that non-biomedical journals won’t be on Pubmed, and they’re absolutely correct! (thanks for catching that, I apologize for being sloppy here). Check out  Web of Science  for a more complete index of science journals. And please feel free to share other resources in the comments!]  Beware of  questionable journals .

As you read, write down  every single word  that you don’t understand. You’re going to have to look them all up (yes, every one. I know it’s a total pain. But you won’t understand the paper if you don’t understand the vocabulary. Scientific words have extremely precise meanings).

Step-by-step instructions for reading a primary research article

1. Begin by reading the introduction, not the abstract.

The abstract is that dense first paragraph at the very beginning of a paper. In fact, that’s often the only part of a paper that many non-scientists read when they’re trying to build a scientific argument. (This is a terrible practice—don’t do it.).  When I’m choosing papers to read, I decide what’s relevant to my interests based on a combination of the title and abstract. But when I’ve got a collection of papers assembled for deep reading, I always read the abstract last. I do this because abstracts contain a succinct summary of the entire paper, and I’m concerned about inadvertently becoming biased by the authors’ interpretation of the results.

2. Identify the BIG QUESTION.

Not “What is this paper about”, but “What problem is this entire field trying to solve?”

This helps you focus on why this research is being done.  Look closely for evidence of agenda-motivated research.

3. Summarize the background in five sentences or less.

Here are some questions to guide you:

What work has been done before in this field to answer the BIG QUESTION? What are the limitations of that work? What, according to the authors, needs to be done next?

The five sentences part is a little arbitrary, but it forces you to be concise and really think about the context of this research. You need to be able to explain why this research has been done in order to understand it.

4.   Identify the SPECIFIC QUESTION(S)

What  exactly  are the authors trying to answer with their research? There may be multiple questions, or just one. Write them down.  If it’s the kind of research that tests one or more null hypotheses, identify it/them.

Not sure what a null hypothesis is? Go read this one  and try to identify the null hypotheses in it. Keep in mind that not every paper will test a null hypothesis.

5. Identify the approach

What are the authors going to do to answer the SPECIFIC QUESTION(S)?

6. Now read the methods section. Draw a diagram for each experiment, showing exactly what the authors did.

I mean  literally  draw it. Include as much detail as you need to fully understand the work.  As an example, here is what I drew to sort out the methods for a paper I read today ( Battaglia et al. 2013: “The first peopling of South America: New evidence from Y-chromosome haplogroup Q” ). This is much less detail than you’d probably need, because it’s a paper in my specialty and I use these methods all the time.  But if you were reading this, and didn’t happen to know what “process data with reduced-median method using Network” means, you’d need to look that up.

Image credit: author

You don’t need to understand the methods in enough detail to replicate the experiment—that’s something reviewers have to do—but you’re not ready to move on to the results until you can explain the basics of the methods to someone else.

7.   Read the results section. Write one or more paragraphs to summarize the results for each experiment, each figure, and each table. Don’t yet try to decide what the results  mean , just write down what they  are.

You’ll find that, particularly in good papers, the majority of the results are summarized in the figures and tables. Pay careful attention to them!  You may also need to go to the Supplementary Online Information file to find some of the results.

 It is at this point where difficulties can arise if statistical tests are employed in the paper and you don’t have enough of a background to understand them. I can’t teach you stats in this post, but  here , and here   are some basic resources to help you.  I STRONGLY advise you to become familiar with them.

Things to pay attention to in the results section:

  • Any time the words “significant” or “non-significant” are used. These have precise statistical meanings. Read more about this  here .
  • If there are graphs, do they have  error bars  on them? For certain types of studies, a lack of confidence intervals is a major red flag.
  • The sample size. Has the study been conducted on 10, or 10,000 people? (For some research purposes, a sample size of 10 is sufficient, but for most studies larger is better).

8. Do the results answer the SPECIFIC QUESTION(S)? What do you think they mean?

Don’t move on until you have thought about this. It’s okay to change your mind in light of the authors’ interpretation—in fact you probably will if you’re still a beginner at this kind of analysis—but it’s a really good habit to start forming your own interpretations before you read those of others.

9. Read the conclusion/discussion/Interpretation section.

What do the authors think the results mean? Do you agree with them? Can you come up with any alternative way of interpreting them? Do the authors identify any weaknesses in their own study? Do you see any that the authors missed? (Don’t assume they’re infallible!) What do they propose to do as a next step? Do you agree with that?

10. Now, go back to the beginning and read the abstract.

Does it match what the authors said in the paper? Does it fit with your interpretation of the paper?

11. FINAL STEP:  (Don’t neglect doing this)  What do other researchers say about this paper?

Who are the (acknowledged or self-proclaimed) experts in this particular field? Do they have criticisms of the study that you haven’t thought of, or do they generally support it?

Here’s a place where I do recommend you use google! But do it last, so you are better prepared to think critically about what other people say.

(12. This step may be optional for you, depending on why you’re reading a particular paper. But for me, it’s critical! I go through the “Literature cited” section to see what other papers the authors cited. This allows me to better identify the important papers in a particular field, see if the authors cited my own papers (KIDDING!….mostly), and find sources of useful ideas or techniques.)

UPDATE: If you would like to see an example of how to read a science paper using this framework, you can find one  here .

I gratefully acknowledge Professors José Bonner and Bill Saxton for teaching me how to critically read and analyze scientific papers using this method. I’m honored to have the chance to pass along what they taught me.

I’ve written a shorter version of this guide for teachers to hand out to their classes. If you’d like a PDF, shoot me an email: jenniferraff (at) utexas (dot) edu. For further comments and additional questions on this guide, please see the Comments Section on  the original post .

This piece originally appeared on the  author’s personal blog  and is reposted with permission.

Featured image credit:  Scientists in a laboratory of the University of La Rioja  by  Urcomunicacion  (Wikimedia CC BY3.0)

Note: This article gives the views of the authors, and not the position of the LSE Impact blog, nor of the London School of Economics. Please review our  Comments Policy  if you have any concerns on posting a comment below.

Jennifer Raff (Indiana University—dual Ph.D. in genetics and bioanthropology) is an assistant professor in the Department of Anthropology, University of Kansas, director and Principal Investigator of the KU Laboratory of Human Population Genomics, and assistant director of KU’s Laboratory of Biological Anthropology. She is also a research affiliate with the University of Texas anthropological genetics laboratory. She is keenly interested in public outreach and scientific literacy, writing about topics in science and pseudoscience for her blog ( violentmetaphors.com ), the Huffington Post, and for the  Social Evolution Forum .

  • Learning Consultations
  • Peer Tutoring
  • Getting Started
  • Peer Education Courses
  • Become a Peer Educator
  • ADHD/LD Support
  • Workshops & Outreach
  • Learning Strategies
  • Manage Time
  • All Resources
  • For Faculty & Staff

SMU Libraries logo

  •   SMU Libraries
  • Scholarship & Research
  • Teaching & Learning
  • Bridwell Library
  • Business Library
  • DeGolyer Library
  • Fondren Library
  • Hamon Arts Library
  • Underwood Law Library
  • Fort Burgwin Library
  • Exhibits & Digital Collections
  • SMU Scholar
  • Special Collections & Archives
  • Connect With Us
  • Research Guides by Subject
  • How Do I . . . ? Guides
  • Find Your Librarian
  • Writing Support

Types of Research Papers: Overview

A research paper is simply a piece of writing that uses outside sources. There are different types of research papers with varying purposes and expectations for sourcing.

While this guide explains those differences broadly, disciplines and assignments vary. Ask your professor for clarification on the purpose and types of appropriate research questions and sources.

Need More Help?

Related guides.

  • Literature Reviews
  • Annotated Bibliographies
  • Starting Your Research

Research and Writing Lab

Need last minute help but didn't book an appointment? Every week we offer online drop-in labs.

Tuesdays 3:00pm - 4:30pm via Zoom @   https://smu.zoom.us/j/92637892352  and in-person, Fondren Red 1st floor (near elevators)

  • Last Updated: Apr 12, 2024 1:00 PM
  • URL: https://guides.smu.edu/researchpapertypes
  • Open access
  • Published: 11 April 2024

Organizing the dissemination and implementation field: who are we, what are we doing, and how should we do it?

  • Gretchen J. R. Buchanan   ORCID: orcid.org/0000-0002-5186-0145 1   na1 ,
  • Lindsey M. Filiatreau 2   na1 &
  • Julia E. Moore 3   na1  

Implementation Science Communications volume  5 , Article number:  38 ( 2024 ) Cite this article

253 Accesses

4 Altmetric

Metrics details

Two decades into its tenure as a field, dissemination and implementation (D&I) scientists have begun a process of self-reflection, illuminating a missed opportunity to bridge the gap between research and practice—one of the field’s foundational objectives. In this paper, we, the authors, assert the research-to-practice gap has persisted, in part due to an inadequate characterization of roles, functions, and processes within D&I. We aim to address this issue, and the rising tension between D&I researchers and practitioners, by proposing a community-centered path forward that is grounded in equity.

We identify key players within the field and characterize their unique roles using the translational science spectrum, a model originally developed in the biomedical sciences to help streamline the research-to-practice process, as a guide. We argue that the full translational science spectrum, from basic science research, or “T0,” to translation to community, or “T4,” readily applies within D&I and that in using this framework to clarify roles, functions, and processes within the field, we can facilitate greater collaboration and respect across the entire D&I research-to-practice continuum. We also highlight distinct opportunities (e.g., changes to D&I scientific conference structures) to increase regular communication and engagement between individuals whose work sits at different points along the D&I translational science spectrum that can accelerate our efforts to close the research-to-practice gap and achieve the field’s foundational objectives.

Peer Review reports

Contributions to the literature

Providing clarity regarding the distinct groups of individuals involved in D&I science and practice from researchers to the communities impacted by the change and outline key roles of these unique sets of actors.

Specifying the range of activities, from theoretical research to applied implementation, involved in D&I science and practice using a translational structure.

Identifying existing gaps (e.g., poor integration of research into existing implementation efforts) that impede attainment of the shared vision of D&I science and practice and propose solutions to these gaps.

Introduction

Though still in its infancy, the field of dissemination and implementation science (D&I) [ 1 , 2 ] is facing challenges related to the growing gap between the science and practice of implementation [ 1 , 3 , 4 ]. D&I is the scientific study of translating research findings and evidence-based interventions into everyday practice; in the current state of the D&I literature, this often means that a practice developed by one group of actors is being implemented into the everyday practice of others [ 5 ]. A premortem by Beidas and colleagues [ 4 ] highlighted several factors stagnating the field, including closure of the evidence-to-practice gap [ 6 , 7 , 8 , 9 ], insufficient impact, and inability to align timelines and priorities with partners [ 1 ]. This commentary aims to establish further clarity regarding who “we” are as a field, what we are doing, and how we can collectively work to achieve shared goals of improved population health in D&I. This refers to the collective “we” of those engaged in D&I work.

In clarifying key components of D&I, important lessons can be drawn from more established fields. For example, when reflecting on disciplines such as mathematics and physics, one notes the emergence of two broad areas of scholarship—theoretical and applied—within these fields. These scholarship areas fill distinct, but important roles within their fields. Here, the authors posit that D&I science could be similarly broken down into theoretical and applied scholarships. In this paper, we, the authors, elaborate on the functions of these differential scholarships, and the functions of professionals working in the large and ever-growing field of implementation practice.

While many have noted D&I aims “to promote the adoption and integration of evidence-based practices, interventions, and policies into routine health care and public health settings to improve the impact on population health,” [ 10 ] specificity in how to achieve this outcome has been elusive. In this article, we propose that the field must first define the actors and audiences across the implementation spectrum and how each group connects with others. Subsequently, the field can strengthen the infrastructures that facilitate these connections. In this article, we aim to address the rising tension between implementation scientists, implementation support practitioners, delivery systems [ 11 ], and communities by proposing a path forward that is community-oriented and grounded in equity, thereby upholding every actor’s place at the D&I table. We draw on principles well-established in the field of translational science to better align D&I towards both improved ideas and real-world impact. We note that our mental model as authors is that success for D&I would be defined as impact at the community or population levels. We recognize this is not the mental model held by all people working in D&I, but believe even for those whose focus is not on population impact, we can collectively work together to achieve these outcomes and impact practice [ 12 ].

Who are we?

To date, much of the discussion around the direction of D&I has been researcher-centric [ 13 ]. To promote greater equity within the discipline (i.e., to reduce disparities in whose voices are heard within the field of D&I), we would like to expand the existing discourse to include the entire spectrum of professionals who work in implementation, including communities, delivery systems, implementation support practitioners, intermediaries, non-implementation science researchers (e.g., interventionists), and applied and theoretical D&I researchers. Including the entire implementation workforce in a description of the field provides opportunities to see where practitioners have not been empowered to exert influence and to change these inequities. While D&I professionals are likely to fill more than one role at a time or during their careers and may hold perspectives that are therefore representative of a number of these D&I actors, we would like to re-center the current conversation within D&I around implementation support practitioners and delivery systems specifically to uphold our commitment to those most directly affected by D&I efforts.

Communities and individuals impacted by the change

Communities and the individuals who comprise them play a critical role in the success or failure of efforts to implement evidence-based or informed programs and practices (EBPs) within a particular setting [ 14 , 15 , 16 , 17 ]. Aligned with this principle, there has been a shifting focus from using community-based to community-led research methods across academic disciplines [ 18 , 19 ]. Funding agencies have also begun to recognize the need for greater community involvement in research, with current directives to engage community partners across the research spectrum [ 20 ]. As suggested by others, strengthening relationships between communities and individuals working at all levels of implementation should remain a priority in closing the evidence-to-practice gap and upholding equity in D;I; indeed, it is essential [ 21 ].

Practitioners—implementation support practitioners and delivery systems

Implementation has been happening for the entirety of human history. While several scientific fields (e.g., political science, medicine) began formally investigating processes of D&I in the mid-to-late twentieth century—thereby laying the foundation for current research in this area— the distinct field of D&I only emerged in the past few decades, prompted by repeatedly observed barriers to the successful implementation of EBPs [ 5 , 22 ].

“Implementation practitioners” are professionals comprised of two distinct groups: implementation support practitioners [ 23 , 24 ] (e.g., administrators, policy-makers) are involved in planning, engagement, co-creation, strategy selection, capacity building, monitoring, and evaluation; delivery systems (e.g., front-line managers at organizations implementing an EBP) are responsible for implementing the actual practices with professionals, organizations, and the public [ 11 ]. Identifying professionals engaged in implementation practice can be difficult as there is inconsistency and terminology; for example, there are over 30 job titles associated with implementation support practitioner roles (see Fig.  1 ). “Delivery systems” are often unaware of the D&I field or their role as end-users. Implementation researchers appropriately identifying and connecting with delivery systems and implementation support practitioners is key to closing the evidence-to-practice gap and improving impact [ 4 ].

figure 1

Professional job titles of individuals working directly in implementation or implementation support as identified through the Center for Implementation (In preparation for an event about the roles of implementation support practitioners, an open call was sent out to members of an online community of professionals supporting implementation. People were asked for their current or previous job titles that included an implementation component.)

Intermediaries

Globally, there are several intermediary organizations serving to translate findings from D&I to support the implementation of EBPs by delivery systems and implementation support practitioners (e.g., the Collaborative for Implementation Practice; Center for Evidence and Implementation in Australia; Impact Center at the University of North Carolina; Center for Effective Services in Ireland; the Nigerian Implementation Science Alliance). These organizations employ implementation support practitioners and bridge the implementation research-to-practice divide by providing training in implementation-related skills and creating tools to support the selection of appropriate implementation strategies. For example, one intermediary has a mini-course providing an introduction to implementation that has enrolled over 10,000 individuals. Millions of research, government, and philanthropic dollars are being invested in these organizations [ 25 , 26 , 27 , 28 ]. As implementation researchers and intermediaries, the authors regularly hear from organizations, communities, and individuals that they struggle to access supports in implementation science to address their needs in implementing evidence The demand for this type of work often outpaces the supply, and researchers and funders alike state a clear need for additional resources linking implementation science and practice [ 29 , 30 , 31 , 32 , 33 ].

Researchers

To better clarify the full spectrum of implementation researchers, researchers whose work is primarily centered on the advancement of implementation ideas (e.g., theory, methods, or framework (TMF) development) are referred to as theoretical implementation scientists and those whose work is primarily centered on the direct use of implementation concepts as a method to achieve better clinical or programmatic outcomes as applied implementation scientists . Scientists may work on both theoretical and applied projects but tend to focus their programs of research in one or the other and may even identify as one or the other.

Non-D&I researchers are also becoming increasingly interested in D&I, as evidenced by the growing number of D&I training institutes globally (e.g., HIV, Infectious Disease and Global Health Implementation Research Institute (HIGH IRI); University College Cork Implementation Science Training Institute; University of Nairobi Implementation Science Fellowship; Training Institute for Dissemination and Implementation Research in Health (TIDIRH)) [ 34 ]. Non-D&I researchers are individuals from distinct substantive areas (e.g., HIV, cancer prevention) who are interested in applying D&I to their work but have limited training in this area. These researchers often aim to draw from the TMFs and evidence from D&I to design, implement, and scale EBPs. They may benefit from increased collaboration with individuals who have worked more squarely in D&I.

What are we doing?

We, the paper’s authors, entered the field of D&I with the goal of bridging the research-to-practice gap to better improve the lives of people in our areas of scholarship (HIV, mental health). Yet, we have found that our substantively distinct bodies of applied D&I research have unfolded in such a way that we are all currently involved in a range of theoretical implementation research. This journey has not been without difficulty—the further we moved from our applied work and what grounded our science, the less impact we felt we were having. While we found theoretical research important, we felt as though our roles and functions within D&I were less clear. This lack of clarity in our professional self-concept ultimately helped us identify that D&I is not monolithic. Through conversation, we found that articulating the spectrum of theoretical to applied D&I helped us regain the clarity we needed to continue advancing our science. We believe these realizations could also be beneficial to other D&I professionals.

Leveraging translational science to find clarity

There is extensive literature on moving research findings into practice [ 35 ], but the translation of D&I knowledge into practice has received much less attention [ 1 ]. Moreover, there is insufficient understanding of which actors are involved at which stages along this spectrum, how each stage contributes to the field, and how these stages, and actors at each of these stages, can connect and achieve shared goals. In Fig.  2 , the authors draw on the translational spectrum to address these limitations. The traditional translational spectrum aims to streamline the “bench to bedside” approach and defines the continuum of basic science (stage T0) to public health science (stage T4) [ 36 ]. D&I science has long been placed in the T3–T4 segments of the traditional translational spectrum [ 36 ]. However, we argue that the full translational spectrum, from T0 through T4, is applicable to D&I. This distinction is often at the core of the tension observed within the field and where our personal struggles in our shifting identities and relationship with D&I research emerged.

figure 2

The translational spectrum applied to implementation science

In the traditional translational spectrum, T0, “pre-clinical research,” includes bench science and aims to define mechanisms, targets, and strategies for intervention on a general level. In D&I, theoretical implementation scientists work on the development of TMFs, and elicitation, description, and modeling of mechanisms. Many of the foundational papers that guide implementation research to date stem from work at this stage [ 37 , 38 , 39 , 40 , 41 , 42 , 43 ]. T1, “translation to humans,” includes Stage 1 clinical trials and proof of concept science and aims to develop new methods of diagnosis, treatment, and prevention in highly controlled settings. In D&I, theoretical and applied implementation researchers focus on translating theoretical constructs (i.e., TMFs) to actual people and developing methods to test these constructs. Examples of this type of research include measurement of implementation domains such as context (e.g., the Organizational Readiness for Change measure) [ 44 ] and implementation outcomes (e.g., the NoMAD measure from Normalization Process Theory) [ 45 ]. T2, “translation to patients,” includes Stages 2–3 clinical trials and aims to develop clinical applications and evidence-based guidelines for a given disease. In D&I, applied implementation researchers focus on identifying implementation constructs relevant to a specific situation, intervention, context, or population where the researchers aim to understand how best to implement. Traditional randomized controlled trial designs are often used in this stage. Individuals working at this stage may test bundled strategies, interrogate the “active ingredients” in strategies [ 46 ], or test strategies in varied contexts.

An interesting phenomenon occurs in the T3–4 range. Acknowledging the contributions of researchers and practitioners, we see a split whereby researchers continue to serve as the primary actors in one branch of the translational spectrum, while practitioners become the primary actors in another branch of the spectrum. T3, “translation to practice,” includes comparative effectiveness trials and clinical outcome studies and aims to evaluate real-world effectiveness. In D&I, implementation support practitioners come into a principal role. Individuals working in this capacity use the results of T0–2 to plan implementation projects, sometimes in the form of quality improvement-type projects. In parallel, T3 applied implementation researchers are primarily monitoring or evaluating implementation projects’ real-world effectiveness; this could involve research using pragmatic or naturalistic methods whereby researchers partner with healthcare delivery systems or organizations to better understand real-world implementation or effectiveness outcomes. T4 involves population-level outcomes research and monitoring improvements in morbidity and mortality to impact policy or system change. In D&I, implementation support practitioners and delivery systems scale EBPs up and out. Implementation researchers working at stage T4 define the implementation workforce, develop surveillance systems, and evaluate the effects of evidence-informed implementation on project successes. Intermediaries are prime partners in this work. Additional work is needed to establish clear evidence about what is and is not working on a broad scale and in what contexts [ 42 , 47 ].

Defining the translational spectrum for D&I facilitates the process of identifying a “home base” for individuals involved in D&I science, thereby improving self-concept clarity and making clear how individuals can foray into upstream and downstream segments to better link their research with that of others. In keeping with findings from workplace self-concept clarity literature [ 48 , 49 ], when we claim our places in the spectrum, we can improve our effectiveness and avoid burnout [ 50 ]. Specifically, we can improve our capacity to clearly generate research questions, identify colleagues, and expand the impact of our work.

How should we do it?

As has been noted by others [ 21 , 51 , 52 ], there is a significant disconnect between individuals working in distinct roles within the field of D&I, particularly between those operating at the two ends of the D&I translational spectrum. By interacting more often and intentionally across the entirety of the D&I process, we as a field could develop significant synergy and produce actionable solutions more quickly to achieve shared goals.

Asking and answering the right question

Fundamental respect for the work of actors at every level of the implementation spectrum, fostered by regular communication, is essential in resolving our identity crises, achieving our shared goals, and upholding equity within the field [ 21 ]. One fundamental way for theoretical implementation scientists to demonstrate respect for implementation practitioners is to ask research questions that implementation practitioners want answered [ 52 ]. Implementation practitioners have critical theoretical questions that arise while implementing programs and policies in their specific contexts. For example, implementation practitioners regularly assess organizational readiness for change before altering or implementing a new program or policy (as recommended in the implementation science literature). Yet when the assessments suggest that sites are not ready to implement the intended change, there is little guidance from implementation science about how to best address this issue. A common suggestion is to prioritize “ready” sites [ 53 ]. This approach is likely to perpetuate existing inequities or disparities, as “ready” sites are often the sites that are least in need of additional resources and supports, and leaves “non-ready” sites with no plan for reaching a sufficient level of readiness. What strategies can increase readiness? Another example involves the need for a more concrete understanding of the effects of adaptation. While the field might agree adaptation is often important to the scale-up and scale-out of EBPs, many adaptation tools [ 54 , 55 ] are designed for researchers as opposed to practitioners looking for guidance in understanding if the adaptations they propose will influence the effectiveness of the original EBP. How can D&I measures be made more accessible for implementation practitioners? These are just two examples of many.

Working with existing implementation efforts

Evaluating existing processes and successes of implementation practitioners can also galvanize efforts, improve impact of D&I, and uphold equity in D&I. Delivery systems are continually implementing “the thing” and have been for years. Connecting with existing implementation efforts and studying the effectiveness of implementation strategies being actively used by delivery systems is critical to supporting the ongoing work of these individuals [ 2 , 21 , 56 ]. In many ways, this can shortcut science more quickly to a clearer understanding of what works when and for whom, and improve the likelihood of establishing sustainable practices and policies that are feasible, acceptable, and appropriate [ 23 , 24 ]. This approach is also consistent with the principles of community-based participatory research, including respect for lived experience and tailoring interventions to the needs of the community [ 57 , 58 ].

Fostering increased communication

Increased communication among actors across the D&I translational spectrum is critical, as previously noted [ 3 , 52 , 59 ]. To again draw from the successes of other fields, the International AIDS Society is a group of over 13,000 members worldwide that “unite(s) scientists, policymakers and activists to galvanize the scientific response, build global solidarity and enhance human dignity for all people living with and affected by HIV” [ 60 ]. The International AIDS Society hosts two conferences that rotate annually with a shifting focus between research and practice. Using this model, which has been repeatedly shown to be highly impactful, individuals working at all stages of the HIV implementation science spectrum can engage in, learn from, and contribute to dialogue with others with distinct perspectives and roles in the discipline, thereby improving equity concerning whose voices are centered and uplifted in global agenda-setting efforts. As such, the field of D&I could benefit from an organization akin to the International AIDS Society and agenda-setting practices and conference structures employed by this Society [ 61 , 62 , 63 ].

Developing tools to directly support real-world D&I

Tools that facilitate the translation of D&I into practice are also critical to achieving shared goals [ 1 ]. Again, the field of D&I can look to adjacent fields to learn how they have successfully scaled. For example, the Institute for Healthcare Improvement (IHI), whose mission is to improve health and healthcare worldwide, has scaled the use of quality improvement methods. Over 30 years, they have worked in 42 countries and have had over 7 million online course enrollments [ 64 ]. Part of IHI’s model has been to develop practical and easy-to-use improvement tools. A critique of implementation science is that existing frameworks are complicated and difficult to use [ 3 , 4 ]. If the field of D&I learned from the success of IHI and developed tools that help professionals operationalize implementation science in practice, it would support the broader use of D&I to improve outcomes.

Aligning funding mechanisms and priorities

Funding agencies should increase requirements and supports for community inclusion and implementation throughout the research process. Researchers currently prioritize funding agency policies and expectations, which may not allow enough time for building sustainable community relationships and co-creation of work. A shift in funding agencies’ research calls and approach to awarding research dollars is necessary to build capacity for long-term academic-community partnerships [ 65 , 66 , 67 ]. Implementation science-related funding calls from the National Institutes of Health, UK Research and Innovation, the Global Alliance for Chronic Diseases, the South African Medical Research Council, and other funding agencies could more intentionally include requirements for this type of work.

Key actions are needed for the field of D&I to self-actualize: (1) Uphold everyone’s place at the implementation table while centering the wants and needs of those most directly affected by implementation efforts; (2) Clarify where on the translational spectrum work is being done by whom and where the gaps in both sufficient volume of work and translation of that work lie; and (3) Facilitate regular communication across the spectrum, from theoretical implementation scientists to implementation practitioners and vice versa. Ideally, this work should be done with researchers and practitioners around the globe. If these three tasks are accomplished, we as a field will be able to reverse the tides and bridge the implementation research-to-practice gap, instead of letting it continue to grow.

Availability of data and materials

Not applicable.

Abbreviations

  • Dissemination and implementation

Human immunodeficiency virus

Acquired immunodeficiency syndrome

Westerlund A, Sundberg L, Nilsen P. Implementation of implementation science knowledge: the research-practice gap paradox. Worldviews Evidence-Based Nurs. 2019;16(5):332–4. https://doi.org/10.1111/wvn.12403 .

Article   Google Scholar  

Metz A, Jensen T, Farley A, Boaz A. Is implementation research out of step with implementation practice? Pathways to effective implementation support over the last decade. Implement Res Pract. 2022;3:263348952211055. https://doi.org/10.1177/26334895221105585 .

Rapport F, Smith J, Hutchinson K, et al. Too much theory and not enough practice? The challenge of implementation science application in healthcare practice. J Eval Clin Pract. 2022;28(6):991–1002. https://doi.org/10.1111/jep.13600 .

Article   PubMed   Google Scholar  

Beidas RS, Dorsey S, Lewis CC, et al. Promises and pitfalls in implementation science from the perspective of US-based researchers: learning from a pre-mortem. Implement Sci. 2022;17(1):1–15. https://doi.org/10.1186/s13012-022-01226-3 .

Eccles MP, Mittman BS. Welcome to implementation science. Implement Sci. 2006;1(1):1–3. https://doi.org/10.1186/1748-5908-1-1 .

Article   PubMed Central   Google Scholar  

Taylor SP, Kowalkowski MA, Beidas RS. Where is the implementation science? An opportunity to apply principles during the COVID-19 pandemic. Clin Infect Dis. 2020:6–8. https://doi.org/10.1093/cid/ciaa622 .

Lyon AR, Comtois KA, Kerns SEU, Landes SJ, Lewis CC. Closing the science–practice gap in implementation before it widens. In: Albers B, Shlonsky A, Mildon R, eds. Implementation Science 3.0. Springer International Publishing; 2020:295–313. https://doi.org/10.1007/978-3-030-03874-8_12 .

Ploeg J, Davies B, Edwards N, Gifford W, Miller PE. Factors influencing best-practice guideline implementation: Lessons learned from administrators, nursing staff, and project leaders. Worldviews Evidence-Based Nurs. 2007;4(4):210–9. https://doi.org/10.1111/j.1741-6787.2007.00106.x .

Bernhardt JM, Mays D, Kreuter MW. Dissemination 2.0: closing the gap between knowledge and practice with new media and marketing. J Health Commun. 2011;16(SUPPL. 1):32–44. https://doi.org/10.1080/10810730.2011.593608 .

National Cancer Institute. Implementation Science. 2020. Accessed 12 Sept 2023. https://cancercontrol.cancer.gov/is/about .

Wandersman A, Duffy J, Flaspohler P, et al. Bridging the gap between prevention research and practice: The interactive systems framework for dissemination and implementation. Am J Community Psychol. 2008;41(3–4):171–81. https://doi.org/10.1007/s10464-008-9174-z .

Boulton R, Sandall J, Sevdalis N. The Cultural Politics of ‘Implementation Science.’ J Med Humanit. 2020;41(3):379–94. https://doi.org/10.1007/s10912-020-09607-9 .

Article   PubMed   PubMed Central   Google Scholar  

Jensen TM, Metz AJ, Farley AB, Disbennett ME. Developing a practice-driven research agenda in implementation science : Perspectives from experienced implementation support practitioners. Published online. 2023. https://doi.org/10.1177/26334895231199063 .

Iwelunmor J, Blackstone S, Veira D, et al. Toward the sustainability of health interventions implemented in sub-Saharan Africa: a systematic review and conceptual framework. Implement Sci. 2016;11(1). https://doi.org/10.1186/s13012-016-0392-8 .

Baptiste S, Manouan A, Garcia P, Etya’ale H, Swan T, Jallow W. Community-led monitoring: when community data drives implementation strategies. Curr HIV/AIDS Rep. 2020;17(5):415–21. https://doi.org/10.1007/s11904-020-00521-2 .

Anderson KA, Dabelko-Schoeny H, Koeuth S, Marx K, Gitlin LN, Gaugler JE. The use of community advisory boards in pragmatic clinical trials: The case of the adult day services plus project. Home Health Care Serv Q. 2021;40(1):16–26. https://doi.org/10.1080/01621424.2020.1816522 .

Ramanadhan S, Davis M, Donaldson ST, Miller E, Minkler M. Participatory Approaches in Dissemination and Implementation Science. In: Brownson RC, Colditz GA, Proctor EK, editors. Dissemination and Implementation Research in Health: Translating Science to Practice. New York: Oxford University Press; 2023. p. 212.

Chapter   Google Scholar  

Ricalde MCA, Annoni J, Bonney R, et al. Understanding the Impact of Equitable Collaborations between Science Institutions and Community-Based Organizations: Improving Science through Community-Led Research. Bioscience. 2022;72(6):585–600. https://doi.org/10.1093/biosci/biac001 .

Fernandez ME, Ten Hoor GA, van Lieshout S, et al. Implementation mapping: using intervention mapping to develop implementation strategies. Front Public Heal. 2019;7(JUN):1–15. https://doi.org/10.3389/fpubh.2019.00158 .

National Institutes of Health. All of Us: Local Community and/or Participant Advisory Boards (C/PABs). 2023. Accessed 13 Sept 2023. https://allofus.nih.gov/about/who-we-are/all-us-community-and-participant-advisory-boards .

Brownson RC, Kumanyika SK, Kreuter MW, Haire-Joshu D. Implementation science should give higher priority to health equity. Implement Sci. 2021;16(1):28. https://doi.org/10.1186/s13012-021-01097-0 .

Morris ZS, Wooding S, Grant J. The answer is 17 years, what is the question: understanding time lags in translational research. J R Soc Med. 2011;104(12):510–20. https://doi.org/10.1258/jrsm.2011.110180 .

Bührmann L, Driessen P, Metz A, et al. Knowledge and attitudes of implementation support practitioners—findings from a systematic integrative review. PLoS One. 2022;17(5 May):1–25. https://doi.org/10.1371/journal.pone.0267533 .

Article   CAS   Google Scholar  

Albers B, Metz A, Burke K. Implementation support practitioners- A proposal for consolidating a diverse evidence base. BMC Health Serv Res. 2020;20(1):1–10. https://doi.org/10.1186/s12913-020-05145-1 .

PCORI dissemination and implementation funding initiatives. Accessed 29 Feb 2024. https://www.pcori.org/impact/putting-evidence-work/pcori-dissemination-and-implementation-funding-initiatives .

United States Agency for International Development. USAID’s Implementation Science Investment. Accessed 29 Feb 2024. https://www.usaid.gov/fact-sheet/usaids-implementation-science-investment .

National Heart, Lung and BI. Implementation Science Branch. Accessed 29 Feb 2024. https://www.nhlbi.nih.gov/about/translation-research-and-implementation-science/implementation-science .

Zurynski Y, Smith CL, Knaggs G, Meulenbroeks I. Funding research translation: how we got here and what to do next. Aust N Z J Public Health. 2021;45(5):420–3. https://doi.org/10.1111/1753-6405.13131 .

Holmes B, Hamilton AB. Three opportunities to boost implementation science at a critical time of need. Heal Published online. 2021. https://doi.org/10.18865/ed.29.S1.77 .

Planning team for the Pathways to Prevention (P2P) Workshop on Achieving Health Equity in Preventive Services and the Office for Disease Prevention portfolio analysis team. We Need More Implementation Science To Improve Health Equity in Clinical Preventive Services. Director’s Messages. https://prevention.nih.gov/about-odp/directors-messages/2022/we-need-more-implementation-science-improve-health-equity-clinical-preventive-services . Published October 14, 2022. Accessed 29 Feb 2024.

Implementation Science Takes Off at Brown. 2023. Accessed 29 Feb 2024. https://psych.med.brown.edu/news/2023-10-02/bridge-growth .

Davis R, D’Lima D. Building capacity in dissemination and implementation science: a systematic review of the academic literature on teaching and training initiatives. Implement Sci. 2020;15(1):97. https://doi.org/10.1186/s13012-020-01051-6 .

Boyce CA, Barfield W, Curry J, et al. Building the next generation of implementation science careers to advance health equity. 2019;29:77–82. https://doi.org/10.18865/ed.29.S1.77 .

Osanjo GO, Oyugi JO, Kibwage IO, et al. Building capacity in implementation science research training at the University of Nairobi. Implement Sci. 2016;11(1):1–9. https://doi.org/10.1186/S13012-016-0395-5 .

Straus SE, Ma JT, Graham I. Defining knowledge translation. Review. 2009;181:165–8. https://doi.org/10.1503/cmaj.081229 .

National Center for Advancing Translational Science. Transforming Translational Science. https://ncats.nih.gov/files/NCATS-factsheet.pdf . Accessed 24 Mar 2020. 2017;Fall.

Proctor E, Silmere H, Raghavan R, et al. Outcomes for implementation research: Conceptual distinctions, measurement challenges, and research agenda. Adm Policy Ment Heal Ment Heal Serv Res. 2011;38(2):65–76. https://doi.org/10.1007/s10488-010-0319-7 .

Damschroder LJ, Aron DC, Keith RE, Kirsh SR, Alexander JA, Lowery JC. Fostering implementation of health services research findings into practice: A consolidated framework for advancing implementation science. Implement Sci. 2009;4(1):1–15. https://doi.org/10.1186/1748-5908-4-50 .

Aarons GA, Hurlburt M, Horwitz SMC. Advancing a conceptual model of evidence-based practice implementation in public service sectors. Adm Policy Ment Heal Ment Heal Serv Res. 2011;38(1):4–23. https://doi.org/10.1007/s10488-010-0327-7 .

Tabak RG, Khoong EC, Chambers DA, Brownson RC. Bridging research and practice: Models for dissemination and implementation research. Am J Prev Med. 2012;43(3):337–50. https://doi.org/10.1016/j.amepre.2012.05.024 .

Walsh-Bailey C, Tsai E, Tabak RG, et al. A scoping review of de-implementation frameworks and models. Implement Sci. 2021;16(1):1–18. https://doi.org/10.1186/s13012-021-01173-5 .

Nilsen P, Bernhardsson S. Context matters in implementation science: A scoping review of determinant frameworks that describe contextual determinants for implementation outcomes. BMC Health Serv Res. 2019;19(1):1–21. https://doi.org/10.1186/s12913-019-4015-3 .

Powell BJ, Waltz TJ, Chinman MJ, et al. A refined compilation of implementation strategies: Results from the Expert Recommendations for Implementing Change (ERIC) project. Implement Sci. 2015;10(1):1–14. https://doi.org/10.1186/s13012-015-0209-1 .

Shea CM, Jacobs SR, Esserman DA, Bruce K, Weiner BJ. Organizational readiness for implementing change: A psychometric assessment of a new measure. Implement Sci. 2014;9(1):1–15. https://doi.org/10.1186/1748-5908-9-7 .

Finch TL, Girling M, May CR, et al. Improving the normalization of complex interventions : part 2 - validation of the NoMAD instrument for assessing implementation work based on normalization process theory ( NPT ). BMC Med Res Methodol. 2018;18(135):1–13.

Google Scholar  

Desveaux L, Nguyen MD, Ivers NM, et al. Snakes and ladders: A qualitative study understanding the active ingredients of social interaction around the use of audit and feedback. Transl Behav Med. 2023;13(5):316–26. https://doi.org/10.1093/tbm/ibac114 .

Edwards N, Barker PM. The importance of context in implementation research. J Acquir Immune Defic Syndr. 2014;67:S157–62. https://doi.org/10.1097/QAI.0000000000000322 .

Gray CE, Mcintyre KP, Mattingly BA, Jr GWL. Interpersonal Relationships and the Self-Concept. Springer International Publishing; 2020.  https://doi.org/10.1007/978-3-030-43747-3 .

Wu P, Liu T, Li Q, Yu X, Liu Z, Tian S. Maintaining the working state of firefighters by utilizing self-concept clarity as a resource. BMC Public Health. 2024;24(1):1–11. https://doi.org/10.1186/s12889-024-17896-1 .

Balundė A, Paradnikė K. Resources linked to work engagement: the role of high performance work practices, employees’ mindfulness, and self-concept clarity. Soc Inq into Well-Being. 2016;2(2):55–62. https://doi.org/10.13165/SIIW-16-2-2-06 .

Harvey G, Rycroft-Malone J, Seers K, et al. Connecting the science and practice of implementation – applying the lens of context to inform study design in implementation research. Front Heal Serv. 2023;3(July):1–15. https://doi.org/10.3389/frhs.2023.1162762 .

Tabak RG, Padek MM, Kerner JF, et al. Dissemination and Implementation Science Training Needs: Insights From Practitioners and Researchers. Am J Prev Med. 2017;52(3):S322–9. https://doi.org/10.1016/j.amepre.2016.10.005 .

Atkins BR, Allred S, Hart D. Philanthropy’s Rural Blind Spot. Stanford Soc Innov Rev. Published online 2021. https://tableau.dsc.umich.edu/t/UM-Public/views/IndexofDeepDisadvantage/CountiesCitiesMap?:isGuestRedirectFromVizportal=y&:embed=y .

Stirman SW, Baumann AA, Miller CJ. The FRAME: An expanded framework for reporting adaptations and modifications to evidence-based interventions. Implement Sci. 2019;14(1):1–10. https://doi.org/10.1186/s13012-019-0898-y .

Miller CJ, Barnett ML, Baumann AA, Gutner CA, Wiltsey-Stirman S. The FRAME-IS: a framework for documenting modifications to implementation strategies in healthcare. Implement Sci. 2021;16(1):36. https://doi.org/10.1186/s13012-021-01105-3 .

Boaz A. Lost in co-production: To enable true collaboration we need to nurture different academic identities . LSE; 2021. p. 1–4. https://blogs.lse.ac.uk/impactofsocialsciences/2021/06/25/lost-in-co-production-to-enable-truecollaboration-we-need-to-%0Anurture-different-academic-identities/ .

Seifer S. Walking the Talk: Achieving the Promise of Authentic Partnerships. Partnersh Perspect. 2007;IV(I):1–12.

Ramanadhan S, Davis MM, Armstrong R, et al. Participatory implementation science to increase the impact of evidence-based cancer prevention and control. Cancer Causes Control. 2018;29(1):363–9. https://doi.org/10.1007/s10552-018-1008-1 .

Shelton RC, Brownson RC. Enhancing Impact: A Call to Action for Equitable Implementation Science. Prev Sci Published online. 2023. https://doi.org/10.1007/s11121-023-01589-z .

International AIDS Society. Home Page. Accessed 2 Oct 2023. https://www.iasociety.org .

Cahn P, McClure C. Beyond the first 25 years: The International AIDS Society and its role in the global response to AIDS. Retrovirology. 2006;3(1):2004–6. https://doi.org/10.1186/1742-4690-3-85 .

Kort R. 5th International AIDS Society Conference on HIV Pathogenesis, treatment and prevention: summary of key research and implications for policy and practice - operations research. J Int AIDS Soc. 2010;13(SUPPL. 1):1–6. https://doi.org/10.1186/1758-2652-13-S1-S5 .

Gayle H, Wainberg MA. Impact of the 16th International Conference on AIDS: can these conferences lead to policy change? Retrovirology. 2007;4:2–3. https://doi.org/10.1186/1742-4690-4-13 .

IHI Marks 30 Years of Quality Improvement in Health Care Worldwide. BusinessWire. https://www.businesswire.com/news/home/20211028005209/en/IHI-Marks-30-Years-of-Quality-Improvement-in-Health-Care-Worldwide . Published 28 Oct 2021.

Elwood WN, Corrigan JG, Morris KA. NIH-Funded CBPR: self-reported community partner and investigator perspectives. J Community Health. 2019;44(4):740–8. https://doi.org/10.1007/s10900-019-00661-6 .

Teufel-Shone NI, Schwartz AL, Hardy LJ, et al. Supporting new community-based participatory research partnerships. Int J Environ Res Public Health. 2019;16(1):1–12. https://doi.org/10.3390/ijerph16010044 .

Minkler M, Blackwell AG, Thompson M, Tamir HB. Community-based participatory research: implications for public health funding. Am J Public Health. 2003;93(8):1210–3. https://doi.org/10.2105/AJPH.93.8.1210 .

Download references

Acknowledgements

We would like to thank Drs. Cory Bradley and Donny Gerke for their contributions in early conceptualization of this paper and to colleagues who took the time to review and provide feedback prior to submission.

GB was supported by the National Institute of Mental Health grant T32MH019960 at Washington University (PI: Leopoldo J. Cabassa) during a portion of manuscript development. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of Mental Health.

Author information

Gretchen J. R. Buchanan, Lindsey M. Filiatreau and Julia Moore are co-first authors.

Authors and Affiliations

Department of Family Medicine and Community Health, Hennepin Healthcare Research Institute, Minneapolis, MN and University of Minnesota Medical School, MN, Minneapolis, USA

Gretchen J. R. Buchanan

Division of Infectious Diseases, School of Medicine, Washington University in St. Louis, MO, St. Louis, USA

Lindsey M. Filiatreau

The Center for Implementation, ON, Toronto, Canada

Julia E. Moore

You can also search for this author in PubMed   Google Scholar

Contributions

GB, LF, and JM equally participated in the conception, drafting, and revising of the manuscript, and they have approved the manuscript as submitted. GB, LF, and JM agree to be personally accountable for their own contributions and to ensure that questions related to the accuracy or integrity of any part of the work, even ones in which the author was not personally involved, are appropriately investigated, resolved, and the resolution documented in the literature.

Corresponding author

Correspondence to Gretchen J. R. Buchanan .

Ethics declarations

Ethics approval and consent to participate, consent for publication, competing interests.

Author JM is the Director of The Center for Implementation and previously led the implementation team at the Knowledge Translation Program, St. Michael’s Hospital. Several examples are drawn from direct experience in these roles. LF and GB declare that they have no competing interests.

Additional information

Publisher’s note.

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Buchanan, G.J.R., Filiatreau, L.M. & Moore, J.E. Organizing the dissemination and implementation field: who are we, what are we doing, and how should we do it?. Implement Sci Commun 5 , 38 (2024). https://doi.org/10.1186/s43058-024-00572-1

Download citation

Received : 10 November 2023

Accepted : 20 March 2024

Published : 11 April 2024

DOI : https://doi.org/10.1186/s43058-024-00572-1

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Implementation science
  • Translational science
  • Translational science spectrum
  • Translational research continuum
  • Evidence to practice gap
  • Implementation support practitioners

Implementation Science Communications

ISSN: 2662-2211

  • Submission enquiries: Access here and click Contact Us
  • General enquiries: [email protected]

parts of a science research paper

Facility for Rare Isotope Beams

At michigan state university, frib researchers lead team to merge nuclear physics experiments and astronomical observations to advance equation-of-state research, world-class particle-accelerator facilities and recent advances in neutron-star observation give physicists a new toolkit for describing nuclear interactions at a wide range of densities..

For most stars, neutron stars and black holes are their final resting places. When a supergiant star runs out of fuel, it expands and then rapidly collapses on itself. This act creates a neutron star—an object denser than our sun crammed into a space 13 to  18 miles wide. In such a heavily condensed stellar environment, most electrons combine with protons to make neutrons, resulting in a dense ball of matter consisting mainly of neutrons. Researchers try to understand the forces that control this process by creating dense matter in the laboratory through colliding neutron-rich nuclei and taking detailed measurements.

A research team—led by William Lynch and Betty Tsang at FRIB—is focused on learning about neutrons in dense environments. Lynch, Tsang, and their collaborators used 20 years of experimental data from accelerator facilities and neutron-star observations to understand how particles interact in nuclear matter under a wide range of densities and pressures. The team wanted to determine how the ratio of neutrons to protons influences nuclear forces in a system. The team recently published its findings in Nature Astronomy .

“In nuclear physics, we are often confined to studying small systems, but we know exactly what particles are in our nuclear systems. Stars provide us an unbelievable opportunity, because they are large systems where nuclear physics plays a vital role, but we do not know for sure what particles are in their interiors,” said Lynch, professor of nuclear physics at FRIB and in the Michigan State University (MSU) Department of Physics and Astronomy. “They are interesting because the density varies greatly within such large systems.  Nuclear forces play a dominant role within them, yet we know comparatively little about that role.” 

When a star with a mass that is 20-30 times that of the sun exhausts its fuel, it cools, collapses, and explodes in a supernova. After this explosion, only the matter in the deepest part of the star’s interior coalesces to form a neutron star. This neutron star has no fuel to burn and over time, it radiates its remaining heat into the surrounding space. Scientists expect that matter in the outer core of a cold neutron star is roughly similar to the matter in atomic nuclei but with three differences: neutron stars are much larger, they are denser in their interiors, and a larger fraction of their nucleons are neutrons. Deep within the inner core of a neutron star, the composition of neutron star matter remains a mystery. 

  “If experiments could provide more guidance about the forces that act in their interiors, we could make better predictions of their interior composition and of phase transitions within them. Neutron stars present a great research opportunity to combine these disciplines,” said Lynch.

Accelerator facilities like FRIB help physicists study how subatomic particles interact under exotic conditions that are more common in neutron stars. When researchers compare these experiments to neutron-star observations, they can calculate the equation of state (EOS) of particles interacting in low-temperature, dense environments. The EOS describes matter in specific conditions, and how its properties change with density. Solving EOS for a wide range of settings helps researchers understand the strong nuclear force’s effects within dense objects, like neutron stars, in the cosmos. It also helps us learn more about neutron stars as they cool.

“This is the first time that we pulled together such a wealth of experimental data to explain the equation of state under these conditions, and this is important,” said Tsang, professor of nuclear science at FRIB. “Previous efforts have used theory to explain the low-density and low-energy end of nuclear matter. We wanted to use all the data we had available to us from our previous experiences with accelerators to obtain a comprehensive equation of state.”   

Researchers seeking the EOS often calculate it at higher temperatures or lower densities. They then draw conclusions for the system across a wider range of conditions. However, physicists have come to understand in recent years that an EOS obtained from an experiment is only relevant for a specific range of densities. As a result, the team needed to pull together data from a variety of accelerator experiments that used different measurements of colliding nuclei to replace those assumptions with data. “In this work, we asked two questions,” said Lynch. “For a given measurement, what density does that measurement probe? After that, we asked what that measurement tells us about the equation of state at that density.”   

In its recent paper, the team combined its own experiments from accelerator facilities in the United States and Japan. It pulled together data from 12 different experimental constraints and three neutron-star observations. The researchers focused on determining the EOS for nuclear matter ranging from half to three times a nuclei’s saturation density—the density found at the core of all stable nuclei. By producing this comprehensive EOS, the team provided new benchmarks for the larger nuclear physics and astrophysics communities to more accurately model interactions of nuclear matter.

The team improved its measurements at intermediate densities that neutron star observations do not provide through experiments at the GSI Helmholtz Centre for Heavy Ion Research in Germany, the RIKEN Nishina Center for Accelerator-Based Science in Japan, and the National Superconducting Cyclotron Laboratory (FRIB’s predecessor). To enable key measurements discussed in this article, their experiments helped fund technical advances in data acquisition for active targets and time projection chambers that are being employed in many other experiments world-wide.   

In running these experiments at FRIB, Tsang and Lynch can continue to interact with MSU students who help advance the research with their own input and innovation. MSU operates FRIB as a scientific user facility for the U.S. Department of Energy Office of Science (DOE-SC), supporting the mission of the DOE-SC Office of Nuclear Physics. FRIB is the only accelerator-based user facility on a university campus as one of 28 DOE-SC user facilities .  Chun Yen Tsang, the first author on the Nature Astronomy  paper, was a graduate student under Betty Tsang during this research and is now a researcher working jointly at Brookhaven National Laboratory and Kent State University. 

“Projects like this one are essential for attracting the brightest students, which ultimately makes these discoveries possible, and provides a steady pipeline to the U.S. workforce in nuclear science,” Tsang said.

The proposed FRIB energy upgrade ( FRIB400 ), supported by the scientific user community in the 2023 Nuclear Science Advisory Committee Long Range Plan , will allow the team to probe at even higher densities in the years to come. FRIB400 will double the reach of FRIB along the neutron dripline into a region relevant for neutron-star crusts and to allow study of extreme, neutron-rich nuclei such as calcium-68. 

Eric Gedenk is a freelance science writer.

Michigan State University operates the Facility for Rare Isotope Beams (FRIB) as a user facility for the U.S. Department of Energy Office of Science (DOE-SC), supporting the mission of the DOE-SC Office of Nuclear Physics. Hosting what is designed to be the most powerful heavy-ion accelerator, FRIB enables scientists to make discoveries about the properties of rare isotopes in order to better understand the physics of nuclei, nuclear astrophysics, fundamental interactions, and applications for society, including in medicine, homeland security, and industry.

The U.S. Department of Energy Office of Science is the single largest supporter of basic research in the physical sciences in the United States and is working to address some of today’s most pressing challenges. For more information, visit energy.gov/science.

Read our research on: Gun Policy | International Conflict | Election 2024

Regions & Countries

Political typology quiz.

Notice: Beginning April 18th community groups will be temporarily unavailable for extended maintenance. Thank you for your understanding and cooperation.

Where do you fit in the political typology?

Are you a faith and flag conservative progressive left or somewhere in between.

parts of a science research paper

Take our quiz to find out which one of our nine political typology groups is your best match, compared with a nationally representative survey of more than 10,000 U.S. adults by Pew Research Center. You may find some of these questions are difficult to answer. That’s OK. In those cases, pick the answer that comes closest to your view, even if it isn’t exactly right.

About Pew Research Center Pew Research Center is a nonpartisan fact tank that informs the public about the issues, attitudes and trends shaping the world. It conducts public opinion polling, demographic research, media content analysis and other empirical social science research. Pew Research Center does not take policy positions. It is a subsidiary of The Pew Charitable Trusts .

IMAGES

  1. Types of research papers

    parts of a science research paper

  2. Parts of a Research Paper

    parts of a science research paper

  3. Seminar: How to Write An Effective Research Paper

    parts of a science research paper

  4. How to Write a Scientific Paper

    parts of a science research paper

  5. Parts of a Research Paper

    parts of a science research paper

  6. Science Research Paper Example

    parts of a science research paper

VIDEO

  1. Research Methods: Extracting the Essentials of a Computer Science Research Paper

  2. What are the sciences?? 😇👍👍

  3. How to Write a Scientific Research Paper

  4. Scoliosis Dialogues: An SRS Podcast

  5. Web of Science research paper reading 6

  6. 3- Section of Introduction: series of "How to write a scientific paper"

COMMENTS

  1. Parts of the paper

    Different sections are needed in different types of scientific papers (lab reports, literature reviews, systematic reviews, methods papers, research papers, etc.). Projects that overlap with the social sciences or humanities may have different requirements. Generally, however, you'll need to include: TITLE. ABSTRACT. INTRODUCTION (Background)

  2. Scientific Writing: Sections of a Paper

    Overview. Typically scientific journal articles have the following sections: References used: Kotsis, S.V. and Chung, K.C. (2010) A Guide for Writing in the Scientific Forum. Plastic and Reconstructive Surgery. 126 (5):1763-71. PubMed ID: 21042135. Van Way, C.W. (2007) Writing a Scientific Paper. Nutrition in Clinical Practice. 22: 663-40.

  3. Scientific Papers

    Scientific Papers. Scientific papers are for sharing your own original research work with other scientists or for reviewing the research conducted by others. As such, they are critical to the ...

  4. 3.2 Components of a scientific paper

    3.2.1 Abstract. The abstract is a short summary (150-200 words or less) of the important points of the paper. It does not generally include background information. There may be a very brief statement of the rationale for conducting the study. It describes what was done, but without details.

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

    Try to give priority to original research articles, rather than reviews. If you want to cite an idea from a paper where the authors already cite another source for the same idea, then you should return to the original article and verify the exactitude of what you are citing, then cite the original authors, not the intermediate paper.

  6. How to Write a Research Paper: Parts of the Paper

    1. The Title. The title should be specific and indicate the theme of the research and what ideas it addresses. Use keywords that help explain your paper's topic to the reader. Try to avoid abbreviations and jargon. Think about keywords that people would use to search for your paper and include them in your title. 2.

  7. How to write a research paper

    Then, writing the paper and getting it ready for submission may take me 3 to 6 months. I like separating the writing into three phases. The results and the methods go first, as this is where I write what was done and how, and what the outcomes were. In a second phase, I tackle the introduction and refine the results section with input from my ...

  8. 1

    A paper is a state-of-the-art communication, a small step in the progress of knowledge. A long-winded Discussion, occupying pages of text with little or no structure, should be avoided. A lengthy Discussion that lacks an internal logical sequence of arguments is disliked by referees, editors and readers.

  9. The Structure of a Scientific Paper

    The readers must know where to quickly find the information they seek, from the cover page to the reference list. This chapter explains the parts of a typical scientific document, how to structure these parts into a well-organized document, and how to write each part to effectively communicate the science. Download to read the full chapter text.

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

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

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

    For student researchers, writing a well-organized research paper is a key step in learning how to express understanding, make critical connections, ... Reader's Guide to A Framework for K-12 Science Education. 7 Table 1 depicts the key skills students should develop as part of the Science as Inquiry Content Standard. ...

  12. 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 ...

  13. PDF The Structure of an Academic Paper

    The thesis is generally the narrowest part and last sentence of the introduction, and conveys your position, the essence of your argument or idea. See our handout on Writing a Thesis Statement for more. The roadmap Not all academic papers include a roadmap, but many do. Usually following the thesis, a roadmap is a

  14. How to Read a Scientific Paper

    The Parts and Uses of Primary Research Articles. Primary research articles are typically broken down into six sections: abstract, introduction, materials and methods, results, discussion, and references. A few journals have slightly different formats due to their space constraints or target audience.

  15. Scientific paper structure: Quick and easy guide

    Scientific paper structure: Key parts. 1. Title and Abstract: Attract the reader's attention. A scientific paper usually starts with two key parts that help attract a reader's attention to your work: the title and abstract. These parts are designed to essentially be the advertisement for your paper.

  16. Research Paper

    Definition: Research Paper is a written document that presents the author's original research, analysis, and interpretation of a specific topic or issue. It is typically based on Empirical Evidence, and may involve qualitative or quantitative research methods, or a combination of both. The purpose of a research paper is to contribute new ...

  17. Writing a Research Paper Introduction

    Table of contents. Step 1: Introduce your topic. Step 2: Describe the background. Step 3: Establish your research problem. Step 4: Specify your objective (s) Step 5: Map out your paper. Research paper introduction examples. Frequently asked questions about the research paper introduction.

  18. Parts of a Research Paper

    Method. This should be the easiest part of the paper to write, as it is a run-down of the exact design and methodology used to perform the research. Obviously, the exact methodology varies depending upon the exact field and type of experiment.. There is a big methodological difference between the apparatus based research of the physical sciences and the methods and observation methods of ...

  19. Ten simple rules for reading a scientific paper

    Scientists write original research papers primarily to present new data that may change or reinforce the collective knowledge of a field. Therefore, the most important parts of this type of scientific paper are the data. Some people like to scrutinize the figures and tables (including legends) before reading any of the "main text": because ...

  20. Main Parts of a Scientific/Technical Paper

    The title of your paper and any needed information about yourself (usually your name and institution). Abstract: A short (usually around 250-400 words) description of the paper. Should include what the purpose of the paper is (including the basic research question/problem), the basic design of your project, and the major findings. Introduction:

  21. Writing a Research Paper for Your Science Fair Project

    When you write your research paper you might want to copy words, pictures, diagrams, or ideas from one of your sources. It is OK to copy such information as long as you reference it with a citation. If the information is a phrase, sentence, or paragraph, then you should also put it in quotation marks. A citation and quotation marks tell the ...

  22. How to read and understand a scientific paper

    1. Begin by reading the introduction, not the abstract. The abstract is that dense first paragraph at the very beginning of a paper. In fact, that's often the only part of a paper that many non-scientists read when they're trying to build a scientific argument. (This is a terrible practice—don't do it.).

  23. PDF A Guide to Developing and Writing Research Papers in Political Science

    The Six Parts of a Research Paper. A research paper in political science typically has 6 parts: (1) Introduction, (2) Literature review, (3) Theory, (4) Research Design, (5) Analysis, and (6) Conclusion/ Discussion. While papers do vary in their construction, that variation usually finds a way to embrace these 6 parts.

  24. Overview

    A research paper is simply a piece of writing that uses outside sources. There are different types of research papers with varying purposes and expectations for sourcing. ... To break down an issue or idea into component parts and present an objective discussion of related perspectives. Answers questions of how or why. Question is focused ...

  25. Organizing the dissemination and implementation field: who are we, what

    Two decades into its tenure as a field, dissemination and implementation (D&I) scientists have begun a process of self-reflection, illuminating a missed opportunity to bridge the gap between research and practice—one of the field's foundational objectives. In this paper, we, the authors, assert the research-to-practice gap has persisted, in part due to an inadequate characterization of ...

  26. FRIB researchers lead team to merge nuclear physics experiments and

    FRIB is the only accelerator-based user facility on a university campus as one of 28 DOE-SC user facilities. Chun Yen Tsang, the first author on the Nature Astronomy paper, was a graduate student under Betty Tsang during this research and is now a researcher working jointly at Brookhaven National Laboratory and Kent State University.

  27. Political Typology Quiz

    About Pew Research Center Pew Research Center is a nonpartisan fact tank that informs the public about the issues, attitudes and trends shaping the world. It conducts public opinion polling, demographic research, media content analysis and other empirical social science research. Pew Research Center does not take policy positions.

  28. Internal Disagreement and Disruptive Technologies

    This paper models the adoption by established firms of internally disruptive technologies in that different parts of an organization stand to lose or gain from adoption. When agents disagree with a decision, they impose costs on the firm.