• Make an Appointment
  • Study Connect
  • Request Workshop

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 .

is powered by WordPress. Read the Sites@Duke Express and , or .

  • 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

Brown University Homepage

Evaluating Information

  • Understanding Primary and Secondary Sources
  • Exploring and Evaluating Popular, Trade, and Scholarly Sources

Reading a Scholarly Article

Common components of original research articles, while you read, reading strategies, reading for citations, further reading, learning objectives.

This page was created to help you:

Identify the different parts of a scholarly article

Efficiently analyze and evaluate scholarly articles for usefulness

This page will focus on reading scholarly articles — published reports on original research in the social sciences, humanities, and STEM fields. Reading and understanding this type of article can be challenging. This guide will help you develop these skills, which can be learned and improved upon with practice.

We will go over:

There are many different types of articles that may be found in scholarly journals and other academic publications. For more, see:

  • Types of Information Sources
Note: Not all articles contain all components.
Title Offers clues to article’s main topic.
Author(s)

Describes who is responsible for this work. May be one person, a group, or an institution. Make note of authors and institutions you see repeatedly during your search process.

Abstract Summarizes article contents and findings; may include methodology.
Keywords

Describe the content in quick words or phrases. Help you place the work in context with other literature. Good for quick reference!

Introduction Summarizes the article’s main idea, thesis, or research question. Should answer the question, "Why this?" Includes background knowledge on the topic and provides information about research motivations, impact, or purpose. 
Literature Review

Places the research in context with prior work. Analyzes important contributions that the author(s) believe are relevant and that the article builds upon to create new knowledge. Sometimes includes a theoretical framework. A good place to look to find additional sources for your research!

Methods (or Methodology)

An explanation of how and why the authors approached the examination of their question and the collection of data. May include information about the limitations of their chosen methodology.

Discussion

An examination of meaning and implications of the research for existing and future exploration.

Figures Graphical representation of findings and other relevant information. Includes charts, graphs, maps, images, tables, etc. Look at figures during your initial scan to determine relevancy and quality.
Conclusion

A synthesis of the findings and importance of the research.

Reading a scholarly article isn’t like reading a novel, website, or newspaper article. It’s likely you won’t read and absorb it from beginning to end, all at once.

Instead, think of scholarly reading as inquiry, i.e., asking a series of questions as you do your research or read for class. Your reading should be guided by your class topic or your own research question or thesis.

For example, as you read, you might ask yourself:

  • What questions does it help to answer, or what topics does it address?
  • Are these relevant or useful to me?
  • Does the article offer a helpful framework for understanding my topic or question (theoretical framework)?
  • Do the authors use interesting or innovative methods to conduct their research that might be relevant to me?
  • Does the article contain references I might consult for further information?

In Practice

Scanning and skimming are essential when reading scholarly articles, especially at the beginning stages of your research or when you have a lot of material in front of you.

Many scholarly articles are organized to help you scan and skim efficiently. The next time you need to read an article, practice scanning the following sections (where available) and skim their contents:

  • The abstract: This summary provides a birds’ eye view of the article contents.
  • The introduction:  What is the topic(s) of the research article? What is its main idea or question?
  • The list of keywords or descriptors
  • Methods: How did the author(s) go about answering their question/collecting their data?
  • Section headings:  Stop and skim those sections you may find relevant.
  • Figures:  Offer lots of information in quick visual format.
  • The conclusion:  What are the findings and/or conclusions of this article?

Mark Up Your Text

Read with purpose.

  • Scanning and skimming with a pen in hand can help to focus your reading.
  • Use color for quick reference. Try highlighters or some sticky notes. Use different colors to represent different topics.
  • Write in the margins, putting down thoughts and questions about the content as you read.
  • Use digital markup features available in eBook platforms or third-party solutions, like Adobe Reader or Hypothes.is.

Categorize Information

Create your own informal system of organization. It doesn’t have to be complicated — start basic, and be sure it works for you.

  • Jot down a few of your own keywords for each article. These keywords may correspond with important topics being addressed in class or in your research paper.  
  • Write keywords on print copies or use the built-in note taking features in reference management tools like Zotero and EndNote.  
  • Your keywords and system of organization may grow more complex the deeper you get into your reading.

Highlight words, terms, phrases, acronyms, etc. that are unfamiliar to you. You can highlight on the text or make a list in a notetaking program.

  • Decide if the term is essential to your understanding of the article or if you can look it up later and keep scanning.

You may scan an article and discover that it isn’t what you thought it was about. Before you close the tab or delete that PDF, consider scanning the article one more time, specifically to look for citations that might be more on-target for your topic.  

You don’t need to look at every citation in the bibliography — you can look to the literature review to identify the core references that relate to your topic. Literature reviews are typically organized by subtopic within a research question or thesis. Find the paragraph or two that are closely aligned with your topic, make note of the author names, then locate those citations in the bibliography or footnote.

See the Find Articles page for what to do next:

  • Find Articles

See the Citation Searching page for more on following a citation trail:

  • Citation Searching
  • Taking notes effectively. [blog post] Raul Pacheco-Vega, PhD
  • How to read an academic paper. [video] UBCiSchool. 2013
  • How to (seriously) read a scientific paper. (2016, March 21). Science | AAAS.
  • How to read a paper. S. Keshav. 2007. SIGCOMM Comput. Commun. Rev. 37, 3 (July 2007), 83–84.

This guide was designed to help you:

  • << Previous: Exploring and Evaluating Popular, Trade, and Scholarly Sources
  • Last Updated: Feb 16, 2024 3:55 PM
  • URL: https://libguides.brown.edu/evaluate

moBUL - Mobile Brown University Library

Brown University Library  |  Providence, RI 02912  |  (401) 863-2165  |  Contact  |  Comments  |  Library Feedback  |  Site Map

Library Intranet

  • UNC Libraries
  • Course Guides
  • ENGL105 - Scholarly Articles 101

How to Read a Scholarly Article

Engl105 - scholarly articles 101: how to read a scholarly article.

  • Peer-review
  • Other Types of Sources
  • Types of Scholarly Articles

Let's Practice!

  • Extend Your Knowledge

Scholarly articles can be very intimidating! They are written by experts for other experts in a field, so they can be filled with technical, confusing jargon. Studies that are measuring something might have long strings of complicated statistics or equations. How are we supposed to get through all of that and figure out what the article is saying? Here are some tips that can help make scholarly articles a little more approachable.

Step 1: Start with the abstract

The purpose of an abstract is to act as a preview for the rest of the article. Usually the abstract will touch on the importance of a topic, the methods used and results found (if a study was involved), and the conclusions the authors reached. All of this should let you know if the article will be a good fit for you and whether you should dedicate more time to reading the entire article or not.

Step 2: Read the introduction and conclusion

The introduction will give you a lot more background information on the topic of the article. The conclusion will tell you what actually happened in the study, as well as the author's interpretation of these results and ideas for future areas of study. Taking the time to read and understand these two sections toward the beginning of your reading process should make the additional sections of the article a little easier to comprehend.

Step 3: Look at any visual data included in the article

Looking at tables, charts, or graphs can often make more sense than reading sentences with long strings of statistics or other numbers. Try to draw some conclusions from this data and then compare your ideas with the author's own conclusions.

Step 4: Go back to the beginning and read the entire article!

You should have a little more knowledge of the ideas the article is discussing and it should be a little easier to understand using that context. As you read, it can often be helpful to take notes, either directly on the article or in your notebook, especially if there are quotes you may want to reference in your project! 

Don't be afraid to Google any terms you don't know and remember that it is okay to take mental breaks if you are feeling overwhelmed!

In the video below, some university professors share their tips for reading scholarly articles:

How to Read a Scholarly Article from NC State University Libraries on Vimeo .

Let's Practice!

I am looking for an article to use as a source for my paper on factors that may impact the mental health of college students. Use the strategies we explored above to quickly skim the article below. Do you think it would be a good source for my paper? Take some time to reflect on why or why not. What section of the paper best helped you decide on your answer?

(you can read the article directly on the site, but it may be a bit easier to navigate the different sections if you download the pdf version).

  • The Prevalence of Depression, Anxiety, and Stress and Their Associated Factors in College Students

Research & Instruction Associate

Profile Photo

  • << Previous: Types of Scholarly Articles
  • Next: Extend Your Knowledge >>
  • Last Updated: Aug 7, 2024 2:00 PM
  • URL: https://guides.lib.unc.edu/scholarly-articles-101

Loading metrics

Open Access

Ten simple rules for reading a scientific paper

* E-mail: [email protected]

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

ORCID logo

  • Maureen A. Carey, 
  • Kevin L. Steiner, 
  • William A. Petri Jr

PLOS

Published: July 30, 2020

  • https://doi.org/10.1371/journal.pcbi.1008032
  • Reader Comments

Table 1

Citation: Carey MA, Steiner KL, Petri WA Jr (2020) Ten simple rules for reading a scientific paper. PLoS Comput Biol 16(7): e1008032. https://doi.org/10.1371/journal.pcbi.1008032

Editor: Scott Markel, Dassault Systemes BIOVIA, UNITED STATES

Copyright: © 2020 Carey et al. This is an open access article distributed under the terms of the Creative Commons Attribution License , which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Funding: 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.

Competing interests: The authors have declared that no competing interests exist.

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.

thumbnail

  • PPT PowerPoint slide
  • PNG larger image
  • TIFF original image

https://doi.org/10.1371/journal.pcbi.1008032.t001

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.

thumbnail

https://doi.org/10.1371/journal.pcbi.1008032.t002

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.

  • 1. Brown E. The Weird Sisters. G. P. Putnam’s Sons; 2011.
  • View Article
  • Google Scholar
  • PubMed/NCBI

Unfortunately we don't fully support your browser. If you have the option to, please upgrade to a newer version or use Mozilla Firefox , Microsoft Edge , Google Chrome , or Safari 14 or newer. If you are unable to, and need support, please send us your feedback .

We'd appreciate your feedback. Tell us what you think! opens in new tab/window

Infographic: How to read a scientific paper

April 5, 2021 | 3 min read

By Natalia Rodriguez

Infographic

Mastering this skill can help you excel at research, peer review – and writing your own papers

Much of a scientist’s work involves reading research papers, whether it’s to stay up to date in their field, advance their scientific understanding, review manuscripts, or gather information for a project proposal or grant application. Because scientific articles are different from other  texts, like novels or newspaper stories, they should be read differently.

Research papers follow the well-known IMRD format — an abstract followed by the  I ntroduction,  M ethods,  R esults and  D iscussion. They have multiple cross references and tables as well as supplementary material, such as data sets, lab protocols and gene sequences. All those characteristics  can make them dense and complex. Being able to effectively understanding them is a matter of practice.

You can use ScienceDirect’s recommendations service to find other articles related to the work you’re reading.  Once you've registered opens in new tab/window , the recommendations engine uses an adaptive algorithm to understand your research interests. It can then find related content from our database of more than 3,800 journals and over 37,000 book titles. The more frequently you sign in, the better it gets to know you, and the more relevant the recommendations you'll receive. Reading a scientific paper should not be done in a linear way (from beginning to end); instead, it should be done strategically and with a critical mindset, questioning your understanding and the findings. Sometimes you will have to go backwards and forwards, take notes and have multiples tabs opened in your browser.

LennyRhine. “ How to Read a Scientific Paper opens in new tab/window ,” Research4Life Training portal

Valerie Matarese, PhD (Ed). “ Usingstrategic, critical reading of research papers to teach scientific writing opens in new tab/window ,” Supporting Research Writing: Rolesand challenges in multilingual settings,” Chandos Publishing, Elsevier (2012)

Allen H. Renear, PhD, and Carole L. Palmer, PhD. " StrategicReading, Ontologies, and the Future of Scientific Publishing opens in new tab/window ," Science (2009).

Angel Borja, PhD. “ 11 steps to structuring a science paper editors will take seriously ,” Elsevier Connect (June 24, 2014)

Mary Purugganan, PhD, and Jan Hewitt, PhD. “ How to Read a Scientific Article opens in new tab/window ,” Cain Project in Engineering andProfessional Communication, Rice University

“How to Read and Review a Scientific Journal Article,”Writing Studio, Duke University

Robert Siegel, PhD. “ ReadingScientific Papers opens in new tab/window ,” Stanford University

Related resources

Elsevier Researcher Academy opens in new tab/window Free e-learning modules developed by global experts; career guidance and advice; research news on our blog.

Research4Life Training Portal opens in new tab/window : A platform with free downloadable resources for researchers. The Authorship Skills section contains 10 modules, including how to read and write scientific papers, intellectual property and web bibliography along with hands-on activity workbooks.

Career Advice portal of Elsevier Connect : Stories include tips for publishing in an international journal, how to succeed in a PhD program, and how to make your mark in the world of science.

Contributor

Natalia Rodriguez

Natalia Rodriguez

  • USC Libraries
  • Research Guides

Organizing Your Social Sciences Research Paper

  • Reading Research Effectively
  • Purpose of Guide
  • Design Flaws to Avoid
  • Independent and Dependent Variables
  • Glossary of Research Terms
  • Narrowing a Topic Idea
  • Broadening a Topic Idea
  • Extending the Timeliness of a Topic Idea
  • Academic Writing Style
  • Applying Critical Thinking
  • Choosing a Title
  • Making an Outline
  • Paragraph Development
  • Research Process Video Series
  • Executive Summary
  • The C.A.R.S. Model
  • Background Information
  • The Research Problem/Question
  • Theoretical Framework
  • Citation Tracking
  • Content Alert Services
  • Evaluating Sources
  • Primary Sources
  • Secondary Sources
  • Tiertiary Sources
  • Scholarly vs. Popular Publications
  • Qualitative Methods
  • Quantitative Methods
  • Insiderness
  • Using Non-Textual Elements
  • Limitations of the Study
  • Common Grammar Mistakes
  • Writing Concisely
  • Avoiding Plagiarism
  • Footnotes or Endnotes?
  • Further Readings
  • Generative AI and Writing
  • USC Libraries Tutorials and Other Guides
  • Bibliography

Reading a Scholarly Article or Research Paper

Identifying a research problem to investigate requires a preliminary search for and critical review of the literature in order to gain an understanding about how scholars have examined a topic. Scholars rarely structure research studies in a way that can be followed like a story; they are complex and detail-intensive and often written in a descriptive and conclusive narrative form. However, in the social and behavioral sciences, journal articles and stand-alone research reports are generally organized in a consistent format that makes it easier to compare and contrast studies and to interpret their contents.

General Reading Strategies

W hen you first read an article or research paper, focus on asking specific questions about each section. This strategy can help with overall comprehension and with understanding how the content relates [or does not relate] to the problem you want to investigate. As you review more and more studies, the process of understanding and critically evaluating the research will become easier because the content of what you review will begin to coalescence around common themes and patterns of analysis. Below are recommendations on how to read each section of a research paper effectively. Note that the sections to read are out of order from how you will find them organized in a journal article or research paper.

1.  Abstract

The abstract summarizes the background, methods, results, discussion, and conclusions of a scholarly article or research paper. Use the abstract to filter out sources that may have appeared useful when you began searching for information but, in reality, are not relevant. Questions to consider when reading the abstract are:

  • Is this study related to my question or area of research?
  • What is this study about and why is it being done ?
  • What is the working hypothesis or underlying thesis?
  • What is the primary finding of the study?
  • Are there words or terminology that I can use to either narrow or broaden the parameters of my search for more information?

2.  Introduction

If, after reading the abstract, you believe the paper may be useful, focus on examining the research problem and identifying the questions the author is trying to address. This information is usually located within the first few paragraphs of the introduction or in the concluding paragraph. Look for information about how and in what way this relates to what you are investigating. In addition to the research problem, the introduction should provide the main argument and theoretical framework of the study and, in the last paragraphs of the introduction, describe what the author(s) intend to accomplish. Questions to consider when reading the introduction include:

  • What is this study trying to prove or disprove?
  • What is the author(s) trying to test or demonstrate?
  • What do we already know about this topic and what gaps does this study try to fill or contribute a new understanding to the research problem?
  • Why should I care about what is being investigated?
  • Will this study tell me anything new related to the research problem I am investigating?

3.  Literature Review

The literature review describes and critically evaluates what is already known about a topic. Read the literature review to obtain a big picture perspective about how the topic has been studied and to begin the process of seeing where your potential study fits within the domain of prior research. Questions to consider when reading the literature review include:

  • W hat other research has been conducted about this topic and what are the main themes that have emerged?
  • What does prior research reveal about what is already known about the topic and what remains to be discovered?
  • What have been the most important past findings about the research problem?
  • How has prior research led the author(s) to conduct this particular study?
  • Is there any prior research that is unique or groundbreaking?
  • Are there any studies I could use as a model for designing and organizing my own study?

4.  Discussion/Conclusion

The discussion and conclusion are usually the last two sections of text in a scholarly article or research report. They reveal how the author(s) interpreted the findings of their research and presented recommendations or courses of action based on those findings. Often in the conclusion, the author(s) highlight recommendations for further research that can be used to develop your own study. Questions to consider when reading the discussion and conclusion sections include:

  • What is the overall meaning of the study and why is this important? [i.e., how have the author(s) addressed the " So What? " question].
  • What do you find to be the most important ways that the findings have been interpreted?
  • What are the weaknesses in their argument?
  • Do you believe conclusions about the significance of the study and its findings are valid?
  • What limitations of the study do the author(s) describe and how might this help formulate my own research?
  • Does the conclusion contain any recommendations for future research?

5.  Methods/Methodology

The methods section describes the materials, techniques, and procedures for gathering information used to examine the research problem. If what you have read so far closely supports your understanding of the topic, then move on to examining how the author(s) gathered information during the research process. Questions to consider when reading the methods section include:

  • Did the study use qualitative [based on interviews, observations, content analysis], quantitative [based on statistical analysis], or a mixed-methods approach to examining the research problem?
  • What was the type of information or data used?
  • Could this method of analysis be repeated and can I adopt the same approach?
  • Is enough information available to repeat the study or should new data be found to expand or improve understanding of the research problem?

6.  Results

After reading the above sections, you should have a clear understanding of the general findings of the study. Therefore, read the results section to identify how key findings were discussed in relation to the research problem. If any non-textual elements [e.g., graphs, charts, tables, etc.] are confusing, focus on the explanations about them in the text. Questions to consider when reading the results section include:

  • W hat did the author(s) find and how did they find it?
  • Does the author(s) highlight any findings as most significant?
  • Are the results presented in a factual and unbiased way?
  • Does the analysis of results in the discussion section agree with how the results are presented?
  • Is all the data present and did the author(s) adequately address gaps?
  • What conclusions do you formulate from this data and does it match with the author's conclusions?

7.  References

The references list the sources used by the author(s) to document what prior research and information was used when conducting the study. After reviewing the article or research paper, use the references to identify additional sources of information on the topic and to examine critically how these sources supported the overall research agenda. Questions to consider when reading the references include:

  • Do the sources cited by the author(s) reflect a diversity of disciplinary viewpoints, i.e., are the sources all from a particular field of study or do the sources reflect multiple areas of study?
  • Are there any unique or interesting sources that could be incorporated into my study?
  • What other authors are respected in this field, i.e., who has multiple works cited or is cited most often by others?
  • What other research should I review to clarify any remaining issues or that I need more information about?

NOTE:   A final strategy in reviewing research is to copy and paste the title of the source [journal article, book, research report] into Google Scholar . If it appears, look for a "cited by" followed by a hyperlinked number [e.g., Cited by 45]. This number indicates how many times the study has been subsequently cited in other, more recently published works. This strategy, known as citation tracking, can be an effective means of expanding your review of pertinent literature based on a study you have found useful and how scholars have cited it. The same strategies described above can be applied to reading articles you find in the list of cited by references.

Reading Tip

Specific Reading Strategies

Effectively reading scholarly research is an acquired skill that involves attention to detail and an ability to comprehend complex ideas, data, and theoretical concepts in a way that applies logically to the research problem you are investigating. Here are some specific reading strategies to consider.

As You are Reading

  • Focus on information that is most relevant to the research problem; skim over the other parts.
  • As noted above, read content out of order! This isn't a novel; you want to start with the spoiler to quickly assess the relevance of the study.
  • Think critically about what you read and seek to build your own arguments; not everything may be entirely valid, examined effectively, or thoroughly investigated.
  • Look up the definitions of unfamiliar words, concepts, or terminology. A good scholarly source is Credo Reference .

Taking notes as you read will save time when you go back to examine your sources. Here are some suggestions:

  • Mark or highlight important text as you read [e.g., you can use the highlight text  feature in a PDF document]
  • Take notes in the margins [e.g., Adobe Reader offers pop-up sticky notes].
  • Highlight important quotations; consider using different colors to differentiate between quotes and other types of important text.
  • Summarize key points about the study at the end of the paper. To save time, these can be in the form of a concise bulleted list of statements [e.g., intro has provides historical background; lit review has important sources; good conclusions].

Write down thoughts that come to mind that may help clarify your understanding of the research problem. Here are some examples of questions to ask yourself:

  • Do I understand all of the terminology and key concepts?
  • Do I understand the parts of this study most relevant to my topic?
  • What specific problem does the research address and why is it important?
  • Are there any issues or perspectives the author(s) did not consider?
  • Do I have any reason to question the validity or reliability of this research?
  • How do the findings relate to my research interests and to other works which I have read?

Adapted from text originally created by Holly Burt, Behavioral Sciences Librarian, USC Libraries, April 2018.

Another Reading Tip

When is it Important to Read the Entire Article or Research Paper

Laubepin argues, "Very few articles in a field are so important that every word needs to be read carefully." However, this implies that some studies are worth reading carefully. As painful and time-consuming as it may seem, there are valid reasons for reading a study from beginning to end. Here are some examples:

  • Studies Published Very Recently .  The author(s) of a recent, well written study will provide a survey of the most important or impactful prior research in the literature review section. This can establish an understanding of how scholars in the past addressed the research problem. In addition, the most recently published sources will highlight what is currently known and what gaps in understanding currently exist about a topic, usually in the form of the need for further research in the conclusion .
  • Surveys of the Research Problem .  Some papers provide a comprehensive analytical overview of the research problem. Reading this type of study can help you understand underlying issues and discover why scholars have chosen to investigate the topic. This is particularly important if the study was published very recently because the author(s) should cite all or most of the key prior research on the topic. Note that, if it is a long-standing problem, there may be studies that specifically review the literature to identify gaps that remain. These studies often include the word "review" in their title [e.g., Hügel, Stephan, and Anna R. Davies. "Public Participation, Engagement, and Climate Change Adaptation: A Review of the Research Literature." Wiley Interdisciplinary Reviews: Climate Change 11 (July-August 2020): https://doi.org/10.1002/ wcc.645].
  • Highly Cited .  If you keep coming across the same citation to a study while you are reviewing the literature, this implies it was foundational in establishing an understanding of the research problem or the study had a significant impact within the literature [either positive or negative]. Carefully reading a highly cited source can help you understand how the topic emerged and how it motivated scholars to further investigate the problem. It also could be a study you need to cite as foundational in your own paper to demonstrate to the reader that you understand the roots of the problem.
  • Historical Overview .  Knowing the historical background of a research problem may not be the focus of your analysis. Nevertheless, carefully reading a study that provides a thorough description and analysis of the history behind an event, issue, or phenomenon can add important context to understanding the topic and what aspect of the problem you may want to examine further.
  • Innovative Methodological Design .  Some studies are significant and should be read in their entirety because the author(s) designed a unique or innovative approach to researching the problem. This may justify reading the entire study because it can motivate you to think creatively about pursuing an alternative or non-traditional approach to examining your topic of interest. These types of studies are generally easy to identify because they are often cited in others works because of their unique approach to investigating the research problem.
  • Cross-disciplinary Approach .  R eviewing studies produced outside of your discipline is an essential component of investigating research problems in the social and behavioral sciences. Consider reading a study that was conducted by author(s) based in a different discipline [e.g., an anthropologist studying political cultures; a study of hiring practices in companies published in a sociology journal]. This approach can generate a new understanding or a unique perspective about the topic . If you are not sure how to search for studies published in a discipline outside of your major or of the course you are taking, contact a librarian for assistance.

Laubepin, Frederique. How to Read (and Understand) a Social Science Journal Article . Inter-University Consortium for Political and Social Research (ISPSR), 2013; Shon, Phillip Chong Ho. How to Read Journal Articles in the Social Sciences: A Very Practical Guide for Students . 2nd edition. Thousand Oaks, CA: Sage, 2015; Lockhart, Tara, and Mary Soliday. "The Critical Place of Reading in Writing Transfer (and Beyond): A Report of Student Experiences." Pedagogy 16 (2016): 23-37; Maguire, Moira, Ann Everitt Reynolds, and Brid Delahunt. "Reading to Be: The Role of Academic Reading in Emergent Academic and Professional Student Identities." Journal of University Teaching and Learning Practice 17 (2020): 5-12.

  • << Previous: 1. Choosing a Research Problem
  • Next: Narrowing a Topic Idea >>
  • Last Updated: Jul 30, 2024 10:20 AM
  • URL: https://libguides.usc.edu/writingguide

info This is a space for the teal alert bar.

notifications This is a space for the yellow alert bar.

National University Library

Research Process

  • Brainstorming
  • Explore Google This link opens in a new window
  • Explore Web Resources
  • Explore Background Information
  • Explore Books
  • Explore Scholarly Articles
  • Narrowing a Topic
  • Primary and Secondary Resources
  • Academic, Popular & Trade Publications
  • Scholarly and Peer-Reviewed Journals
  • Grey Literature
  • Clinical Trials
  • Evidence Based Treatment
  • Scholarly Research
  • Database Research Log
  • Search Limits
  • Keyword Searching
  • Boolean Operators
  • Phrase Searching
  • Truncation & Wildcard Symbols
  • Proximity Searching
  • Field Codes
  • Subject Terms and Database Thesauri

Reading a Scientific Article

  • Website Evaluation
  • Article Keywords and Subject Terms
  • Cited References
  • Citing Articles
  • Related Results
  • Search Within Publication
  • Database Alerts & RSS Feeds
  • Personal Database Accounts
  • Persistent URLs
  • Literature Gap and Future Research
  • Web of Knowledge
  • Annual Reviews
  • Systematic Reviews & Meta-Analyses
  • Finding Seminal Works
  • Exhausting the Literature
  • Finding Dissertations
  • Researching Theoretical Frameworks
  • Research Methodology & Design
  • Tests and Measurements
  • Organizing Research & Citations This link opens in a new window
  • Picking Where to Publish
  • Bibliometrics
  • Learn the Library This link opens in a new window

Library Tutorial

  • Reading a Scholarly Article Tutorial This interactive tutorial provides practice reading a scholarly or scientific article.

Additional Resources

  • Anatomy of a Scholarly Article
  • How to Read (and Understand) a Social Science Journal Article
  • How to Read a Scientific Paper
  • How to Read a Scientific Paper Interactive Tutorial
  • How to Read Scientific Literature (YouTube Video)
  • How to read and understand a scientific paper: a guide for non-scientists See it in action - How to read a vaccine safety study: an example

General Dictionaries

  • The American Heritage Dictionary of the English Language
  • The American Heritage Student Science Dictionary
  • The Chambers Dictionary
  • Dictionary.com
  • The Free Dictionary
  • Merriam-Webster's Collegiate Dictionary
  • Merriam-Webster Online
  • The Penguin English Dictionary
  • The Science Dictionary

Attempting to read a scientific or scholarly research article for the first time may seem overwhelming and confusing. This guide details how to read a scientific article step-by-step. First, you should not approach a scientific article like a textbook— reading from beginning to end of the chapter or book without pause for reflection or criticism. Additionally, it is highly recommended that you highlight and take notes as you move through the article. Taking notes will keep you focused on the task at hand and help you work towards comprehension of the entire article.

  • Skim the article. This should only take you a few minutes. You are not trying to comprehend the entire article at this point, but just get a basic overview. You don’t have to read in order; the discussion/conclusions will help you to determine if the article is relevant to your research. You might then continue on to the Introduction. Pay attention to the structure of the article, headings, and figures.  
  • Grasp the vocabulary. Begin to go through the article and highlight words and phrases you do not understand. Some words or phrases you may be able to get an understanding from the context in which it is used, but for others you may need the assistance of a medical or scientific dictionary. Subject-specific dictionaries available through our Library databases and online are listed below.  
  • The abstract gives a quick overview of the article. It will usually contain four pieces of information: purpose or rationale of study (why they did it); methodology (how they did it); results (what they found); conclusion (what it means). Begin by reading the abstract to make sure this is what you are looking for and that it will be worth your time and effort.   
  • The introduction gives background information about the topic and sets out specific questions to be addressed by the authors. You can skim through the introduction if you are already familiar with the paper’s topic.  
  • The methods section gives technical details of how the experiments were carried out and serves as a “how-to” manual if you wanted to replicate the same experiments as the authors. This is another section you may want to only skim unless you wish to identify the methods used by the researchers or if you intend to replicate the research yourself.  
  • The results are the meat of the scientific article and contain all of the data from the experiments. You should spend time looking at all the graphs, pictures, and tables as these figures will contain most of the data.  
  • Lastly, the discussion is the authors’ opportunity to give their opinions. Keep in mind that the discussions are the authors’ interpretations and not necessarily facts. It is still a good place for you to get ideas about what kind of research questions are still unanswered in the field and what types of questions you might want your own research project to tackle. (See the Future Research Section of the Research Process for more information).  
  •   Read the bibliography/references section. Reading the references or works cited may lead you to other useful resources. You might also get a better understanding of the basic terminology, main concepts, major researchers, and basic terminology in the area you are researching.  
  • Have I taken time to understand all the terminology?
  • Am I spending too much time on the less important parts of this article?
  • Do I have any reason to question the credibility of this research?
  • What specific problem does the research address and why is it important?
  • How do these results relate to my research interests or to other works which I have read?  
  • Read the article a second time in chronological order. Reading the article a second time will reinforce your overall understanding. You may even start to make connections to other articles that you have read on this topic.

Reading a Scholarly Article & Finding Definitions Webinar

This webinar presents effective techniques for reading and understanding a scholarly article, as well as locating definitions related to your research topic.

Subject-Specific Dictionaries

  • Health Sciences
  • Marriage & Family Science
  • Research Methods
  • Social Work

Book jacket for The AMA Dictionary of Business and Management

Was this resource helpful?

  • << Previous: Subject Terms and Database Thesauri
  • Next: Evaluating Information >>
  • Last Updated: Aug 8, 2024 6:27 PM
  • URL: https://resources.nu.edu/researchprocess

National University

© Copyright 2024 National University. All Rights Reserved.

Privacy Policy | Consumer Information

Banner

How to Read Scholarly Articles: Strategies for Reading

  • What is Scholarly?
  • The Anatomy of a Scholarly Article
  • Strategies for Reading
  • Where to Find Scholarly Articles

How to Read

reading research articles

The fact is, these scholars are experts in their field writing for other experts. They are using specialized language that can be difficult for someone new to understand. So, you can sit down with an article and start reading, but you may become discouraged pretty quickly.

The tips below are to help you read scholarly articles STRATEGICALLY . These tips can help you approach a scholarly text for easier reading and better understanding. 

1. Abstract

Read the Abstract first.  The Abstract will preview the entire article, makes it easier to judge whether it is relevant.

For the Sciences:

Titles can only tell you so much about the content of the article. The Abstract acts as a preview for the entire article, including the methods and results. By reading the Abstract first, you can get a better idea of what the article is actually about, if it relates to what you are researching, and whether it is worth your time to read the rest of it.

For the Humanities:

  • Articles in the Arts and Humanities do not always include an Abstract, and if they do, it might just be the first paragraph of the introduction. If not included, move onto the Introduction. Make sure to skim through the section headings, if they are there. This will give you an idea of the organization of the article as well as a general idea of themes.

2. Intro and Conclusion

Next, read the Introduction and Conclusion.  Learn more about the topic of study and what the authors found out in the process.

Applies for both sciences and humanities:

  • These two sections give you the background information you need for the topic of the article as well as what happened in the study. The introduction also includes info about previous studies/papers that relate to the current one, which gives you, the reader, a context. By reading the conclusion you see whether the study answered the original research question and what the authors see as the next steps in the scholarship.

Literature review : An overview of previous scholarship on the present topic. Gives both author and reader a context for where the article falls in the literature. Likely to be a separate section within the introduction or right after it.

Take a look at the tables, charts and graphs.

Get a better idea of the results of the research or analytical study. 

  • Closely look at the visual representations of the data. See what conclusions you come to and make note of them. When you read through the entire article, compare your conclusions to what the authors saw in their results and data.
  • Usually, there is no numeric data that the authors present in their results. However, there might be other visual representations of what the scholars are studying. For example, reproductions of art pieces, or excerpts from primary sources or literary pieces. These are worth looking at to see the materials being studied.

4. Read the Whole Thing

Read it! (For real this time.)

reading research articles

Now that you have pre-read some of the article and are sure it relates to your research topic, read the whole thing. It still might not be easy, but it will not be as hard as if you were reading it with no context.

Some more tips about reading:

  • Summarize sections or paragraphs
  • Keep a subject dictionary, your textbook glossary or the Internet/Wikipedia close by. If you come across any unfamiliar terms, you can quickly look them up.
  • Keep track of the citation information of the articles you do read and want to use in your paper or assignment. This will make life a lot easier at the end of the project. 
  • Reading in the Humanities and Social Sciences Short guide from Trent University with helpful questions to think about to get the most of reading scholarly articles.
  • Reading and Annotating Slideshow from the University of New England. Introduces methods of reading and how to annotate materials.
  • A Guide to Reading and Analysing Academic Articles A guide from Yukon College, discussing the steps for effective reading of academic articles
  • How to Read a Psychology Article Class website from UIC. Includes specific information about how to read articles in Psychology
  • Anatomy of a Scholarly Article Tutorial on how to identify parts of a scholarly article. Created by North Carolina State University.
  • << Previous: The Anatomy of a Scholarly Article
  • Next: Where to Find Scholarly Articles >>
  • Last Updated: May 8, 2024 9:59 AM
  • URL: https://researchguides.ccc.edu/hw/scholarlyarticles
  • Your Science & Health Librarians
  • How To Find Articles with Databases
  • Video Learning
  • Artificial Intelligence Tools
  • Industry Reports
  • How To Evaluate Articles
  • Search Tips, General
  • Develop a Research Question
  • How To Read A Scientific Paper
  • How To Interpret Data
  • How To Write A Scientific Paper
  • Teaching Materials
  • Systematic & Evideced-Based Reviews
  • Get More Help

Useful Sources

  • How to (Seriously) Read a Scientific Paper
  • How to Read a Scientific Article
  • Infographic: How to Read a Scientific Paper

Reading a Scientific Paper

Reading a scientific paper can seem like a daunting task. However, learning how to properly read a scholarly article can make the process much easier! Understanding the different parts of a scientific article can help the reader to understand the material. 

  • The title of the article can give the reader a lot of information about its contents, such as the topic, major ideas, and participants. 
  • Abstracts help to summarize the article and give the reader a preview of the material they are about to read. The abstract is very important and should be read with care. 

Introduction

  • What is the article's purpose being stated in the introduction?
  • Why would this article be of interest to experts in the field?
  • What is already known, or not known, about this topic? 
  • What specifically is the hypothesis? If one is not given, what are the expectations of the author?
  • Having these questions in mind when reading the introduction can help the reader gain an understanding of the article as a whole. A good research article will answer these questions in the introduction and be consistent with their explanation throughout the rest of the article. 
  • What are the specific methods used by the researcher?
  • Does the researcher provide a coherent and viable plan for their experiment?
  • Has the author missed any variables that could effect the results of their findings?
  • How do the methods in this article compare with similar articles?
  • Ex: they are correlated and support the hypothesis, they contradict they hypothesis, ect. 
  • If there are differences from the hypothesis, what differences did the researcher find?
  • Are the findings described in an unbiased way?
  • Is there new information presented that wasn't known before?
  • Is the researcher unbiased in their presentation?
  • Ex: More research needs to be done, the findings show a solution to a known problem, etc.
  • What suggestions are made about future research? If no suggestions are made, should there be?
  • The conclusion points out the important findings from the experiment or research. Occasionally, it will incorporated into the discussion section of the paper. 

General Tips

  • Fully comprehending a scientific article will most likely take more than one read. Don't be discouraged if you don't understand everything the first time, reading scientific papers is a skill that is developed with practice. 
  • Start with the broad and then to the specific. Begin by understanding the topic of the article before trying to dig through all the fine points the author is making. 
  • Always read the tables, charts, and figures. These will give a visual clue to the methods and results sections of the paper and help you to understand the data. The author put these in the paper for a reason, don't dismiss their importance. 
  • Don't be afraid to ask questions or look up definitions. If you do not understand a term or concept, do not be afraid to ask for help or look up an explanation. 
  • << Previous: Develop a Research Question
  • Next: How To Interpret Data >>
  • Last Updated: Jul 26, 2024 11:56 AM
  • URL: https://guides.libraries.indiana.edu/STEM

Social media

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

Additional resources

Featured databases.

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

IU Libraries

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

Organizing Your Social Sciences Research Paper: Reading Research Effectively

  • Purpose of Guide
  • Writing a Research Proposal
  • Design Flaws to Avoid
  • Independent and Dependent Variables
  • Narrowing a Topic Idea
  • Broadening a Topic Idea
  • The Research Problem/Question
  • Academic Writing Style
  • Choosing a Title
  • Making an Outline
  • Paragraph Development
  • The C.A.R.S. Model
  • Background Information
  • Theoretical Framework
  • Citation Tracking
  • Evaluating Sources
  • Reading Research Effectively
  • Primary Sources
  • Secondary Sources
  • What Is Scholarly vs. Popular?
  • Is it Peer-Reviewed?
  • Qualitative Methods
  • Quantitative Methods
  • Common Grammar Mistakes
  • Writing Concisely
  • Avoiding Plagiarism [linked guide]
  • Annotated Bibliography
  • Grading Someone Else's Paper

Reading a Scholarly Article or Research Paper

Reading Research Publications Effectively

It's easy to feel overwhelmed and frustrated when first reading a scholarly article or research paper. The text is dense and complex and often includes abstract or convoluted language . In addition, the terminology may be confusing or applied in a way that is unfamiliar. To help overcome these challenges w hen you first read an article or research paper, focus on asking specific questions about each section. This strategy can help with overall comprehension and understanding how the content relates [or does not relate] to the research problem you are investigating. This approach will also help identify key themes as you read additional studies on the same topic. As you review more and more studies about your topic, the process of understanding and critically evaluating the research will become easier because the content of what you review will begin to coalescence around common themes and patterns of analysis.

Think about the following in this general order:

1.  Read the Abstract

An abstract summarizes the basic content of a scholarly article or research paper. Questions to consider when reading the abstract are: What is this article about? What is the working hypothesis or thesis? Is this related to my question or area of research? The abstract can be used to help filter out sources that may have appeared useful when you began searching for information but, in reality, are not relevant.

2.  Identify the Research Problem and Underlying Questions? 

If, after reading the abstract, you believe the paper may be useful, focus on examining the research problem and identifying the questions the author is trying to address. Look for information that is relevant to your research problem and make note of how and in what way this information relates to what you are investigating.

3.  Read the Introduction and Discussion/Conclusion

The introduction provides the main argument and theoretical framework of the article. Questions to consider for the introduction include what do we already know about this topic and what is left to discover? What other research has been conducted about this topic? How is this research unique? Will this study tell me anything new related to the research problem I am investigating?

Questions to ask yourself while reading the discussion and conclusion sections include what does the study mean and why is it important? What are the weaknesses in their argument? Does the conclusion contain any recommendations for future research and do you believe conclusions about the significance of the study and its findings are valid?

4.  Read about the Methods/Methodology

If what you have read so far closely relates to your research problem, then move on to reading about how the author(s) gathered information for their research. Questions to consider include how did the author do the research? Was it a qualitative, quantitative, or mixed-methods project? What data is the study based on? Could I repeat their work and is all the information available to repeat the study?

5.  Read about the Results and Analysis

Next, read the outcome the research and how it was discussed and analyzed. If any non-textual elements [e.g., graphs, charts, tables, etc.] are confusing, focus on the explanations about them in the text. Questions to consider are what did the author find and how did they find it? Are the results presented in a factual and unbiased way? Does their analysis of results agree with the data presented? Is all the data present? What conclusions do you formulate from this data and does it match with the author's conclusions?

6.  Review the References

The list of references, or works cited, shows you the basis of prior research used by the author(s) to support their study. The references can be an effective way to identify additional sources of information on the topic. Questions to ask include what other research studies should I review? What other authors are respected in this field, i.e., who is cited most often by others? What other research should be explored to learn about issues I am unclear or need more information about?

Reading Tips

Preparing to Read a Scholarly Article or Research Paper for the First Time

Reading scholarly publications effectively is an acquired skill that involves attention to detail and the ability to comprehend complex ideas, data, and concepts in a way that applies logically to the research problem you are investigating. Here are some strategies to consider.

While You are Reading

  • Focus on information in the publication that is most relevant to the research problem
  • Think critically about what you read and seek to build your own arguments; not everything is 100% true or examined effectively
  • Read out of order! This isn't a novel or movie; you want to start with the spoiler
  • Look up the definitions of words you don't know as you read

There are any number of ways to take notes as you read, but use the method that you feel most comfortable with. Taking notes as you read will save time when you go back to examine your sources. Below are some suggestions:

  • Print the article and highlight, circle, and/or underline text as you read [or, you can use the highlight text   feature in a PDF document]
  • Take notes in the margins [Adobe Reader offers pop-up sticky notes]
  • Focus on highlighting important quotes; consider using a different color to differentiate between quotes and other types of text you want to return to when writing
  • Quickly summarize the main or key points at the end of the paper

As you read, write down questions that come to mind that relate to or may clarify your research problem. Here are a few questions that might be helpful:

  • Have I taken time to understand all the terminology?
  • Am I spending too much time on the less important parts of this article?
  • Are there any issues that the authors did not consider?
  • Do I have any reason to question the credibility of this research?
  • What specific problem does the research address and why is it important?
  • How do these results relate to my research interests or to other works which I have read?

Adapted from text originally created by Holly Burt, USC Libraries, April 2018. Thank you, Holly!

  • << Previous: Evaluating Sources
  • Next: Primary Sources >>
  • Last Updated: Sep 8, 2023 12:19 PM
  • URL: https://guides.library.txstate.edu/socialscienceresearch

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
  • HHS Author Manuscripts

Logo of nihpa

The Science of Reading and Its Educational Implications

Mark s. seidenberg.

Department of Psychology, University of Wisconsin-Madison

Research in cognitive science and neuroscience has made enormous progress toward understanding skilled reading, the acquisition of reading skill, the brain bases of reading, the causes of developmental reading impairments and how such impairments can be treated. My question is: if the science is so good, why do so many people read so poorly? I mainly focus on the United States, which fares poorly on cross-national comparisons of literacy, with about 25-30% of the population exhibiting literacy skills that are low by standard metrics. I consider three possible contributing factors, all of which turn on issues concerning the relationships between written and spoken language. They are: the fact that English has a deep alphabetic orthography; how reading is taught; and the impact of linguistic variability as manifested in the Black-White “achievement gap”. I conclude that there are opportunities to increase literacy levels by making better use of what we have learned about reading and language, but also institutional obstacles and understudied issues for which more evidence is badly needed.

Is there an area in cognitive science and cognitive neuroscience that has been more successful than the study of reading? Let's not underestimate the amount that has been learned. We, the community of scientists who study reading (including my colleagues Perfetti and Treiman, whose own research is described in accompanying articles) understand the basic mechanisms that support skilled reading, how reading skill is acquired, and the proximal causes of reading impairments.

We understand the fundamental problem facing the beginning reader: how to relate a new code, a written script, to an existing code, spoken language. We know which behaviors of 4 year old pre-readers are strong predictors of later reading ability, how children make the transition from pre-reader to reader, and the obstacles that many encounter. We know what distinguishes good and poor readers, younger and older skilled readers, “typical” readers from those who are atypical because of either constitutional factors (such as a hearing or learning impairment) or environmental ones (for example, poor schooling or poverty).

We know how basic skills that provide the child's entry into reading relate to other types of knowledge and capacities that support comprehending texts of increasing variety and difficulty. We understand that some aspects of reading are universal (because people's brains are essentially alike) and that some are not (because of differences among writing systems and the languages they represent).

Neuroimaging studies have been successful in identifying the main brain circuits involved in reading and the anomalous ways they develop in dyslexics, and several probable causes of such impairments. We have computational models that specify the mechanisms that underlie basic reading skills, how children acquire them, and how differences in experience (with spoken language and reading) and individual differences (in learning and memory capacities, motivation and other factors) result in varied reading outcomes. This vast research base has led to the development of intervention and remediation methods that can reliably help many children who need it. Researchers disagree about many details—it's science, not the Ten Commandments—but there is remarkable consensus about the basic theory of how reading works and the causes of reading successes and failures (for reviews, see Rayner et al., 2001; Pennington, 2006 ; Morris et al., 2010; Gabrieli, 2009 ; Pugh et al., 2012 ).

My question, then, is this: if the science is so advanced, why do so many people read so poorly? In America not long ago we had a “Sputnik moment,” occasioned by the release of the results of the 2009 round of the PISA cross-national assessments of the academic performance of 15 year olds (OECD PISA, 2009). As in previous years, US performance was close to the average for the 34 OECD countries. However, this round was the first to include data from Shanghai and Singapore, which along with Korea, Hong Kong, and Japan, scored higher than the US. These findings received far more attention than the fact that for many years the US has scored lower than countries such as Canada, Australia, Finland and New Zealand on the PISA exercise. The president, the secretary of education, and the commentariat (e.g., Finn, 2009) all treated the results as evidence of a crisis in American education that called for immediate action. But 2010 was not 1957 and so the second “Sputnik moment” passed quickly, rapidly dropping out of public discourse ( Fig. 1 ).

An external file that holds a picture, illustration, etc.
Object name is nihms503624f1.jpg

Number of times the phrase “sputnik moment” was uttered on CNN, an American cable news network over a two year period. The Larege spike followed the release of results from the 2009 PISA assessment and coincided with President Obama's 2011 State of the Union address.

Although the PISA results made the news, there is plenty of in-house data about the literacy problem in the US, posted on the Department of Education's web site ( http://ies.ed.gov ). The National Assessment of Adult Literacy (2003) found that about 93 million adults read at “basic” or “below basic” levels. At these levels, a person might be able to find the listing for a television program on cable TV, but not understand the instructions and warnings that come with their blood pressure medication ( Lesgold & Welch-Ross, 2012 ). Results on the NAEP (“the nation's report card”) document the origins of low literacy in the performance of 4 th and 8 th graders ( http://nces.ed.gov/nationsreportcard ). Like everything else about education in the US, this assessment exercise has been the focus of controversy, with different stakeholders spinning the data in different ways. People who emphasize how well American education is doing point to the finding that since 1992, when the modern form of the NAEP was introduced, between 59-67% of 4 th graders and 69-76% of 8 th graders scored at the “basic” level or higher. People who think we should be doing better—I am in this camp—can point to the fact that 66-71% of 4 th graders and 66-71% of 8 th graders scored at “basic” or “below basic” levels. Put another way, there are far too many children scoring in the lowest tier (about a third of the 4 th graders and a quarter of the 8 th graders are “below basic”), and far too few in the highest (6-8% of the 4 th graders and a mere 3% of the 8 th graders are “advanced” readers by this measure). The American polity is in a test-happy phase and so there is much other data about who can read and how well than can be reviewed here. It is fair to say that assessments of adults and children consistently indicate that large numbers of individuals in the US read poorly, and that this has been true for many years.

Low literacy's consequences for the affected individuals and for society are vast, as we all know. It creates serious challenges to fully participating in the workforce, managing your own health care, and advancing your children's education. Looking at these facts, and knowing something about how reading works, I've asked myself whether our science has anything to contribute to improving literacy outcomes in this country and others. It might not. Literacy failure could be due to factors well outside the boundaries of this science: poverty, for example. Poverty has many sequelae, including higher infant mortality rate, atypical brain development, shorter life span, worse health and health care, higher crime and incarceration rates, lower educational achievement, higher dropout rates, poorer schools with less experienced teachers and toward the bottom of a list that could go on much longer, poor reading ( General Accounting Office, 2007 ). Surely reducing poverty would have a bigger impact on literacy than anything inspired by our research. Any person with a politically-acceptable plan to substantially reduce or eliminate poverty should step forward immediately.

If poverty were all that mattered, this article could end here. However, the relationship between socioeconomic status and reading achievement is not simple. It is difficult to isolate effects of SES (itself a complex construct; Duncan & Magnuson, 2005 ) from the many other factors with which it is correlated. Nonetheless, data from a variety of sources suggest that there is much about observed literacy outcomes that SES does not explain. I return to this issue in the final section of this article in the context of the Black-White “achievement gap”, where the confound between SES and achievement is of particular concern. Here I merely want to cite some representative findings suggesting that although poverty has enormous impact, it is not the whole story.

The PISA assessments provide a wealth of data (so to speak) about the relationship between national wealth and reading performance. The 2009 data set includes multiple measures related to a country's economic health for a core group of 34 OECD countries. The main findings are quite interesting. 1 In brief, two economic factors, the country's gross domestic product (GDP) and amount spent on education, are only weakly related to reading performance. The proportion of socio-economically disadvantaged students in each country has a bigger impact, with higher proportions associated with lower scores. However, the US does not score poorly because lower income students are overrepresented; in fact, the US clusters with many OECD countries on this measure, and reading scores in this group vary widely. It is also of interest that parents' education level is a much stronger predictor than economic indicators across countries. Of course, it takes more complex analyses to identify relations among such factors and their relative contributions. Nonetheless, even these descriptive data indicate that both SES and other factors are important determinants of outcomes.

The NAEP assessment also includes information regarding the moderating effects of a variety of factors, including race/ethnicity, gender, and eligibility for subsidized or free school lunch, a common (though rough) proxy for SES. 2 Again there are strong indications that both SES and other factors affect outcomes. There is a large, consistent effect of SES as indexed by the subsidized lunch proxy in every year of testing. However, there are similar results for other factors, such as gender. Females have scored significantly higher than males in every year of NAEP testing. Females also scored significantly higher than males in every participating country/municipality in the 2009 PISA assessment. (The US had one of the smaller gender gaps, whereas Finland, a perennial high-scoring country, had one of the largest.) The gender differences within and across countries may be related in some complex manner to SES, but the consistency of the effect across countries with widely varying economic profiles suggests that SES is not the main determinant.

Insofar as the US does not seem likely to substantially reduce or eliminate poverty any time soon and SES is not the only factor affecting reading outcomes, this article will not end here. To restate the question: what are the main causes of reading failures in the US (and perhaps other countries where similar conditions exist) and does reading science have anything to contribute to substantially reducing them, the considerable impact of poverty notwithstanding? I'll consider this question by examining three quite different kinds of factors often thought to be relevant to literacy outcomes in the US.

Blame English?

One possibility is that a certain number of people are doomed to fail to learn to read well because of intrinsic properties of English. The child's initial task is to learn how the written code relates to the spoken language they already know. The writing system is alphabetic, and we tell beginning readers that letters correspond to sounds, but then we teach them early reading vocabulary that includes HAVE, GIVE, SAID, SOME, WAS, WERE, IS, ME, ONE, WHO, SCHOOL, and many other words with atypical spelling-sound correspondences. These inconsistencies are a much commented-upon property of English. Other alphabetic writing systems are indeed more consistent at this level of analysis, many of them conforming (to a very high degree but not perfectly) to the principle that each symbol in the writing system (a “grapheme” consisting of one or more letters) correspond to a single unit (a phoneme) in the spoken language. English is said to be a “deep” orthography, whereas Italian, German, Russian, Finnish, Korean, Serbo-Croatian, and many other alphabets are “shallow” ( Katz & Frost, 1992 ). Written English is obviously a workable system but the learning curve is steep and a greater proportion of individuals may be left behind than if the writing system were shallow.

This hypothesis is contradicted by the consistently high reading achievement in countries such as Canada, New Zealand, Australia, and Singapore where English is also the main language of instruction (Quebec exception duly noted). Although these results suggest that written English is not the whole problem, perhaps performance would be even higher in these countries (and the US) if it were not so peculiar. These cross-national findings are correlational of course. What is needed is more direct evidence as to whether it is easier to learn to read in shallow alphabetic orthographies, holding other factors aside to the extent possible.

Researchers in many countries have attempted to address this question. By now it is quite clear that it is easier to learn to read words and nonwords aloud in shallow alphabetic orthographies compared to English (for reviews, see Aro & Wimmer, 2003 , and several chapters in Snowling & Hulme, 2005, and in Joshi & Aaron, 2006). The advantage for shallow orthographies has been observed in Italian, Spanish, German, French, Finnish, Serbian, Turkish and other languages. “Learning to read in Albanian” is “A skill easily acquired” according to Hoxhallari et al. (2004) because the alphabet is so shallow. Children know the full set of spelling-sound correspondences for Finnish, which has a shallow orthography, by the time formal schooling commences (at age 7 following a compulsory year of preschool). 3 Tested on their skill at reading words and nonwords aloud, children in Wales learning to read in Welsh (which has a shallow alphabetic script) outperform children from the same area who are learning to read in English ( Hanley et al., 2004 ). The Welsh studies permitted comparisons that excluded many potentially confounding socioeconomic and cultural factors. Such studies suggest in short, that shallow is easier. Share (2008) argues that theories of reading have been led astray because of overreliance on studies of English, an “outlier” among writing systems. Perhaps there would be higher literacy achievement in the US if the writing system were more like Finnish or Albanian.

I don't think so. For one thing, this comparative research on reading acquisition makes the mistake of equating the task of reading words and nonwords aloud with “reading” (as in Aro & Wimmer, 2004, Spencer & Hanley, 2003, and many other studies). Children are sometimes called upon to read aloud, in classrooms and in experiments; reading aloud provides overt evidence about the child's knowledge of words and the opportunity to provide explicit feedback (e.g., corrections of mispronunciations). Because of the nature of the writing system, the child's ability to name words and nonwords aloud in English is a major step in reading acquisition. The task has also provided a domain in which to explore statistical learning procedures (Harm & Seidenberg, 1999) that are relevant to language acquisition, visual cognition, and much else. The goal of reading, however, is comprehension. Reading aloud is much more strongly related to comprehension in English than in shallow orthographies (see, e.g., Lindgren, deRenzi, & Richman, 1985 ). In shallow orthographies, reading aloud can be achieved without comprehending what is being said, indeed without knowing the language. I know this to be true because I proved it at my Bar Mitzvah. Modern Hebrew can be written with or without vowels. With the vowels included, the writing system is shallow: words have simple and consistent spelling-sound correspondences, which can be learned rapidly, comprehension not required. Fortuitously, Hebrew is a good “Bar Mitzvah language” ( Seidenberg, 2011 ), as are Finnish, Albanian, Welsh, Italian and other shallow alphabetic orthographies.

One would not want to confuse barking at print with reading comprehension, however. Phil Gough would not have. According to his “simple view of reading” ( Hoover & Gough, 1990 ), children's reading comprehension is a function of decoding skills (recognizing letters, relating print to sound) and knowledge of spoken language (vocabulary and grammar). These skills are dissociable. If the writing system is sufficiently shallow, a person can learn to read aloud without comprehension (my Hebrew). Conversely, a person can know a spoken language quite well without being able to read it (as is true of most 5 year olds who speak English). Among clinicians and researchers, there is a move to reserve the term “dyslexia” for a developmental reading impairment that interferes with acquiring basic print-related skills, especially ability to relate print to sound, independent of spoken language comprehension ( Snowling & Hulme, 2011 ). Other children acquire adequate decoding skills but comprehend texts poorly; these children are also “poor readers,” but a different diagnostic category is needed because their poor reading comprehension is secondary to deficiencies in spoken language.

Granted that it is easy to learn to decode in shallow orthographies, does this confer a comprehension advantage as well? Few studies have closely examined reading aloud, reading comprehension, and spoken language abilities in the same children, although there are some interesting leads. The studies of children learning to read in Welsh and English yielded an interesting tradeoff: whereas the Welsh children performed much better at reading common words and simple nonwords aloud, the English children scored higher when tested on comprehension. As Hanley et al. (2004) noted, “This result suggests that a transparent orthography does not confer any advantages as far as reading comprehension is concerned. As comprehension is clearly the goal of reading, this finding is potentially reassuring for teachers of English” (p. 1408). Why the English children exhibited better comprehension with poorer reading aloud cannot be determined with certainty from these studies. The comparison between Welsh-learning and English-learning children is not entirely clean because the sociolinguistic context is such that English is the dominant language. The Welsh-learning children therefore have substantial knowledge of English (and are bilingual to some degree), whereas the English-learning children have much less knowledge of Welsh and are essentially monolingual. What is clear is that ability to read aloud may say little about the child's reading comprehension.

Durgunoğlu (2006) reached a similar conclusion from extensive studies of reading in Turkish. Turkish has a shallow orthography and a complex, highly productive agglutinating morphology. Summarizing, she noted that “Phonological awareness and decoding develop rapidly in both young and adult readers of Turkish because of the transparent orthography and the special characteristics of phonology and morphology. However, reading comprehension is still a problem.” (2006, p. 226). In her experiments, children's comprehension lagged substantially behind their ability to pronounce words aloud, which she attributes to properties of the spoken language, specifically that complex morphological system, which takes native speakers many years to learn. Whereas comprehension develops more rapidly than production in learning a first language, shallow orthographies create the opposite effect: production—reading aloud—can advance more rapidly than comprehension.

Even within English, accuracy in reading aloud and reading comprehension frequently decouple. Skilled readers are able to read and comprehend many words they mispronounce. Here are some I collected from students and colleagues—words they did not know how to pronounce or systematically mispronounced for many years:

EgregiousQuay
PiquantHegemony
SuaveAutomata
RapportChaos
CoitusFacade
ClitorisEnnui
EpitomeSleight
SegueUranus

All true. A graduate student who spent a portion of his youth immersed in the computer game Chaos: The Battle of Wizards didn't realize until much later that it was connected to the spoken form /ke i -αs/. The two pronunciations of URANUS seem to be in free variation in the US. People can be more adept at engaging in coitus than pronouncing it. A personal example: the Seidenberg and McClelland (1989) model learned to pronounce QUAY as /kwe i / because the training lexicon was created by hand and that is how I thought it was pronounced (it was corrected in later models). As a land-locked kid growing up on the south side of Chicago, I knew the word from reading but not speech. These cases show that a person can know the meaning of a written word but lack secure knowledge of the pronunciation. People frequently generate erroneous pronunciations that they would not have heard in spoken language. If words like these were used in a reading-aloud experiment, even adult, highly skilled readers of English would perform more poorly than Welsh or Turkish subjects.

Nation and Cocksey (2009) found that 7 year old English-speaking children often know the meanings of words they incorrectly read aloud. Familiarity with the spoken form of a word (as indexed by auditory lexical decision performance) was related to accuracy in reading it aloud, especially for words with irregular spelling-sound correspondences. Across subjects, 521 words were read aloud incorrectly; the correct definition was provided for 328 of them (63%). Of course the fact that ability to comprehend and pronounce words can dissociate should have been obvious from the mere existence of severely hearing impaired deaf individuals who do not receive oral training, do not know the pronunciations of words, but are nonetheless good readers. (The fact that it difficult to become a skilled reader under these conditions is an important but separate issue; Goldin-Meadow & Mayberry, 2001 ).

In summary, reading aloud is not a good index of reading comprehension or a basis for evaluating “ease of learning to read” different writing systems. We should therefore be skeptical of claims that it is easy to learn to read in shallow orthographies, and of the corollary belief that English is especially difficult. Over the past 20 years or so, researchers in many countries have correctly recognized the importance of obtaining data about reading in languages other than English, but attempted to correct the imbalance by replicating studies that had been conducted in English using reading aloud, a task that is more closely related to comprehension in that language precisely because of orthographic idiosyncrasies they were trying to surmount.

The relationship between writing systems and spoken languages

The orthographic depth hypothesis is an example of a very interesting idea that drew attention to an important issue (differences in how writing systems represent phonology and their potential impact on reading) and stimulated an enormous amount of research but turned out to be wrong. The hypothesis narrowly focused on the computation of phonology from print. Given the dependence of reading on spoken language, it seemed to follow that writing systems for which it was easier to compute phonology, the shallow ones, would also be easier to comprehend, other factors being equal. This prediction did not turn out to be correct because other factors are manifestly unequal. Looking across languages and writing systems, it can be seen that the properties of writing systems are related to properties of the languages they represent, in particular the complexity of the language's inflectional morphology. Inflectional morphology is an especially important component of language because it is an interface system conveying information about words and the syntactic structures in which they participate, and a major source of typological variation. Languages such as Welsh and Turkish have shallow writing systems but they are morphologically complex, marking properties such as case, number and gender. English and the Sinitic languages (Mandarin, Taiwanese, Cantonese, et al.) exhibit the opposite pattern: the writing systems are deep but their inflectional systems are simple. Looking at English, Gough had observed that early reading comprehension is a function of knowledge of print and knowledge of spoken language. With a cross-linguistic perspective it becomes clear that the two components are not independent. What has to be learned about print depends on properties of the writing system, which bear a non-arbitrary relationship to spoken language typology.

I have attempted to unify these broad cross-linguistic tendencies under the concept of “grapholinguistic equilibrium” ( Seidenberg, 2011 ). The writing systems that have survived support comprehension about equally well. A writing system's capacity to support comprehension can be thought of as a constant that is maintained via trade-offs between orthographic complexity (“depth”, number and complexity of symbols, etc.) and spoken language complexity (particularly morphosyntactic). For languages such as Welsh or Turkish, the spelling-sound correspondences are easily learned, but the morphology is not. These conditions allow children to accurately read aloud sentences that they would not be able to produce or fully comprehend given their still-developing knowledge of the spoken language. Written English is deep but the inflectional system is trivial and little impediment to comprehension. Under these conditions, children easily produce and comprehend sentences that they cannot accurately read aloud.

A deep orthography would be highly dysfunctional, possibly unlearnable, in languages with complex morphosyntax. To illustrate consider Serbo-Croatian. Classic studies focused on its highly consistent spelling-sound correspondences, so different from those in English ( Katz & Frost, 1992 ). My colleagues and I have been more interested in its inflectional system ( Mirković et al., 2004 , 2011 ), which is also very different from English. The system is unquestionably complex. Both nouns and verbs are inflected and there are inflections for number, gender, case, and tense. The inflections are not independent: Number on nouns, for example, depends on case and gender. The inflections are not discrete beads on a string, either: the system is fusional, such that a single suffix encodes multiple inflections. Then there is an additional wrinkle: the realization of an inflection depends on phonological properties of the root to which it is attached ( Table 1 ). The base form SAVETNIK (masculine, “advisor”) is zero-inflected. The final consonant K [/k/] is not retained throughout the inflectional paradigm, changing to C [/ts/] and Č [/tʃ/]. The inflection –E is used for both the vocative singular and the accusative plural; in the former, it is preceded by Č, in the latter by K. It is clear even from this sliver of the language that there is a lot to learn.

WordCase and numberInflection
SAVETNIKNominative singularZero inflected
SAVETNICINominative plural-I
SAVETNIKAGenitive singular-A
SAVETNIČEVocative singular-E
SAVETNIKEAccusative plural-E

Note: All forms are masculine gender. K = /k/ as in Kevin, Č = /t?/ as in “church,” C = /ts/ as in “pizza.”

Now imagine trying to read this language in a script that is more like English, with single letters that represent multiple vowels (e.g., DOSE, LOSE, POSE). Then toss in a few random consonants with multiple pronunciations, such as C (as in CAP and CENT), G (GOAT, GIN), and Y (YOUNG, EDGY). The complexity of the inflectional system is already high. Adding ambiguity in the pronunciations of written letters would increase it enormously. The proper form of an inflection depends on the pronunciation of the previous consonant, but now the pronunciation of the letter representing that consonant will sometimes also depend on context (as with the ambiguous English letters). There would be further penalties if mastery of a complex morphological system requires formal instruction that itself involves reading.

It would take quantitative analyses or simulation models to determine the effects of additional orthographic indeterminacy and establish when the system would become intractable for human learners. The historical fact that languages with complex morphological systems have shallow orthographies is itself suggestive of pressures to maintain this equilibrium, however. Indeed, many times the alignment of language and writing system has been achieved with active intervention, as with the Armenian alphabet in the fifth century, Hangul in 15th century Korea, and Serbo-Croatian in the 19 th century (see Daniels & Bright, 1996 ).

In summary, there is no free orthographic lunch. The child does gain entry into reading more quickly if the associations between units in the written and spoken languages are simple and consistent. However, learning to read aloud in shallow writing systems is a bit like learning to play the violin in the Suzuki method. Both allow the child to rapidly begin performing with relatively little instruction. A four year old's performance of the “Twinkle Variations” may well be the musical equivalent of barking at print. Being able to pronounce words aloud is a helpful skill to possess if your task is to learn a complex, quasiregular morphological system over a many-year period that extends into formal schooling. But, there is little evidence that precocious knowledge of spelling-sound correspondences confers a comprehension advantage, or that the irregularities in written English present an especial burden. 4

What About How Reading is Taught?

American educators have never been able to settle on how to teach children to read. The issue has been debated since Horace Mann was head of the Massachusetts Board of Education in the 1840s. Mann described letters as “skeleton-shaped, bloodless, ghostly apparitions” and encouraged teaching children to read whole words at a time—a lesson that “would be like an excursion to the fields of Elysium” compared to other practices. Mann's tone—authoritative assertion coupled with contempt for other views—is characteristic of much of the subsequent 150 years of debate. 5

How much of the literacy problem in America is due to the way reading has been taught? Everyone knows about the “reading wars” of the past 30 years–the debate over “phonics” and “whole language” approaches. The 2000s saw the emergence of a compromise called “Balanced Literacy,” said to incorporate the best aspects of the two approaches. “Balanced literacy” is a Treaty of Versailles solution that allowed educators to declare the increasingly troublesome “wars” over without having seriously addressed the underlying causes of the strife. The issues are complex, controversial, and ongoing. Here I want to briefly examine some basic considerations, from the perspective of a scientist who studies how reading works, which suggest that how reading is taught is indeed a significant part of the literacy problem in the US and other countries. There are three main points: (a) Contemporary reading science has had very little impact on educational practice mainly because of a two-cultures problem separating science and education; (b) This disconnection has been harmful. Current practices rest on outdated assumptions about reading and development that make learning to read harder than it needs to be, a sure way to leave many children behind; (c) Connecting the science to educational practice would be beneficial but is extremely difficult to achieve. The current environment limits the amount of collaborative work at the all-important translational interface. In the US, the conflicting and often strongly entrenched interests of various stakeholders—educators, politicians, scientists, taxpayers, labor organizations, parent groups—make it hard to achieve meaningful change within the existing institutional structure of public education.

My comments about the culture of education (by which I mean beliefs and attitudes about how children learn and the functions of schooling, particularly with respect to reading) may seem harsh to readers who are not close to the issues. Many people will naturally assume that although scientists and educators may have different views, both have much to contribute and the path to greater progress is through cooperation. Every academic is aware of the importance of interdisciplinary work and of the challenges involved in communicating across disciplines. We also know that the successful creation of cross-disciplinary bridges can have transformative effects, sometimes leading to the emergence of new fields that are much more than the sum of the disciplinary parts. Such a transformation is needed in education and I hope it can be achieved. The question is how. It may be hard for people who are unfamiliar with the landscape to appreciate just how difficult the challenges are. As someone who has been immersed in these issues for many years I have struggled with finding ways to have a positive impact, and that is reflected in the material that follows (see also Seidenberg, 2012).

You may believe, as I usually do, that the collegial and politically-astute approach is to assume that well-intentioned individuals can transcend their differences in the service of a shared goal. Disciplinary barriers only exist as long as we allow them to. We can all do better jobs communicating what we do and what we've learned. Bridges are built on a foundation of mutual respect for individuals and diverse viewpoints. People are doing the best they can; neither side knows everything. I fully support creative bridge-building and have engaged in it myself, but I have come to question whether good intentions and greater effort can be any more effective going forward than they have been in the past. These positive and sincere impulses might have a better chance of succeeding if there were better understanding of the deep differences between the cultures of science and education, which are manifested in their discordant approaches to reading (see also Seidenberg, forthcoming ).

It is important to note that there is plenty of good science relevant being conducted within schools of education, often in departments such as Educational Psychology; however, it is isolated from programs focused on professional training and the development of curricula and instructional practices. My comments on the culture of education focus on the training-and-practices side. I should also stress that my concerns are not about teachers, but rather about what teachers are taught (about child development in general and reading in particular) and about how curricula and instructional practices are created and evaluated. I am not challenging anyone's integrity, commitment, motivation, effort, sincerity, or intelligence. But I am challenging some deeply-held beliefs that have guided educational policies and practices for many years. I would expect this to be discomfiting for many people, but also recognizable as relevant to their deep commitments to helping students learn.

Finally, I must acknowledge that my treatment of these issues is incomplete, given this article's length limitations. Below I mainly characterize the current situation rather than how it arose. The resistance to the reading science of the recent past also needs to be considered in a historical context, which includes earlier attempts to base educational practices on the science of the moment. It also needs to be considered in light of other challenges to educators' traditional control over educational policies and practices (including federal intervention via legislation such as No Child Left Behind, and powerful new educational philanthropies; Ravitch, 2011 ). I provide this broader context in Seidenberg (forthcoming) .

The Two Cultures of Science and Education

Learning to read is an educational issue, historically the purview of educators, specifically schools of education. The history of education in the US has been extensively documented, mainly from the perspective of educators themselves (e.g., Ravitch, 2000 ; Cremin, 1988 ). Popping up a level, one sees that science and education occupy different territory in the intellectual world (literally so on many university campuses). The result is that people who are studying the same thing—how children learn to read, for example—can nonetheless have little contact. The cultures of education and science are radically different: they have different goals and values, ways of training new practitioners, criteria for evaluating progress. The two cultures also communicate their research at separate conferences sponsored by parallel professional organizations attended by different audiences, and publish their work in different journals. There are publishers that target one audience and not the other. These cross-cultural differences, like many others, are difficult to bridge.

Psychologists have been studying reading since the 19 th century and educators have had an approach-avoidance conflict about it ever since. Education as a discipline embraced a few theorists with roots in modern psychology—Dewey, Vygotsky, Piaget, and Bruner among others—whose work underlies the deeply entrenched “constructivist” approach in education ( Tobias & Duffy, 2009 ). There is deep skepticism about the relevance of empirical studies that utilize the tools of modern experimental cognitive and developmental psychology, whether in laboratories or classroom settings (e.g., Coles, 2000 ); however, it co-exists with a readiness to appropriate findings that are consistent with existing beliefs and practices. The special role of science—to find out, to the best of our ability, what is true, letting the implications fall where they may—is subverted if people selectively attend to the findings they find congenial: it transforms research studies into another form of anecdote. Educators also use our research as a source of novel findings that feed the relentless demand for educational innovation. Often this means getting far too carried away far too rapidly with findings that are interesting and new but also not solidly established or understood.

These conflicting attitudes about science and education are at the heart of controversies about reading instruction. What I'll call the Modern Synthesis about learning to read, reading skill, and the relationship between reading and language emerged from work conducted since the 1970s, beginning with Gibson and Levin (1978) , Liberman et al. (1977) , Gough (e.g., Gough & Hillinger, 1980 ; Hoover & Gough, 1990 ), Stanovich (1980) , and others. Almost all of this research was conducted by scientists working outside traditional departments and schools of education. The empirical findings underlying the Modern Synthesis were summarized in several white papers commissioned by various agencies ( Adams, 1990 ; Snow et al., 1998 ; National Reading Panel, 2000 ; Snow, 2002 ; Lonigan & Shanahan, 2009 ). This research called into question basic assumptions underlying how reading is taught and what teachers are taught about reading and development—most importantly the idea that the way that children acquire a first, spoken language provides a good model for learning to read—and yet it has had little subsequent impact on them. The conflicts between scientific and educational approaches to reading continue, centered on three issues.

1. Deciding what is true

One of the major cross-cultural differences concerns attitudes about evidence. There is a movement to encourage evidence-based practices in education, analogous to the ones in medicine and clinical psychology (see http://ies.ed.gov/ncee/wwc ). The effort founders, however, if the stakeholders do not agree on what counts as evidence, or who should decide. Many educators are dismissive of attempts to examine reading from a scientific perspective, which is seen as sterile and reductive, intrinsically incapable of capturing the ineffable character of the learning moment, or the chemistry of a successful classroom ( Coles, 2000 ). Education as a discipline has placed much higher value on observation and hard-earned classroom experience. This division was apparent in reactions to the NRP report (2000) . The panel reviewed the scientific literature relevant to learning to read, having established explicit a priori criteria for what kinds of studies would be considered. Those criteria excluded studies that educators value: mainly, observational, quasi-ethnographic studies of individual schools, teachers, classrooms, and children that do not attempt to conform to basic principles of experimental design or data analysis (see, e.g., Barton & Hamilton, 1998; Rasinski et al., 2011 ). The report was therefore of little interest, to many educators except as evidence for a scientistic bias at odds with the educational establishment's core values ( Krashen, 2001 ). 6

From the perspective of modern studies of cognition, educators' confidence in the reliability of their own observations and experiences in classroom settings is baffling. If teachers really could figure out how reading works and children learn just by observation and experience, there wouldn't be a literacy problem or debates about best practices. But what we can learn about reading this way is limited. Most of what we do when we read is subconscious: we are aware of the result—whether we understood a text or not, whether we found the information we were seeking. Neither teachers nor scientists can directly observe children's mental and neural processes; what can be intuited about them based on classroom experience is limited, and intuitions often conflict. Introspection and systematic personal observation were the main methodologies used by the founders of modern psychology ( Boring, 1953 ), but discovery of their limitations led to the adoption of less observer-dependent methods. The limitations are even greater than the early psychologists could have known. What people observe depends on what they believe (see Cox et al., 2004 , for a striking illustration). Inferences based on observation are subject to deep-seated biases that required Nobel-prize caliber research to uncover ( Kahneman, 2011 ). The limitations of personal observation and experience are among the reasons why we conduct this other, scientific, kind of research: to understand components of reading that would otherwise be hidden from view and to do it in an objective, independently verifiable way. A folk psychology about how we read based on intuition and observation does not become any more reliable when elevated to educational principle, but that is the modern history of educational theorizing about reading.

2. The socio-cultural approach

The Modern Synthesis developed out of research that examined reading within the broader context of research on human language and cognition and their neural and computational bases. Within education, a much more influential approach has emphasized the socio-cultural aspects of literacy, particularly the status of reading in different cultural, linguistic, and socio-economic subgroups (e.g., Gee, 1997 ; Au, 1998 ; Scribner & Cole, 1981 ; Moje & Luke, 2009 ). The approach emphasizes attitudes toward reading within such groups; the varied purposes for which people read in different contexts defined by situation, culture, language, or SES; the relevance of different reading-related activities to learners in these contexts; and how socio-cultural factors affect a child's motivation to learn to read and which classroom practices will be successful.

Much of what is assumed within the socio-cultural approach seems true enough, at an informal level: reading isn't a unitary task; how we read depends on what we are reading and for what purpose; in developing a curriculum it would be wise to take into account the cultural and socio-economic context, including different attitudes toward reading and differences in experiences and opportunities outside the classroom that can greatly affect children's progress. These factors are likely to have a strong impact on the child's motivation to read, a very significant factor that reading scientists have largely ignored.

The socio-cultural research addresses important issues; they are deeply implicated in the “achievement gaps” discussed in the next section. The problem is that socio-cultural paradigm is positioned as an alternative to studies of the types of knowledge and processing mechanisms that underlie reading and how they are acquired, rather than addressing complementary issues. The tension between these approaches furnished the subtext for the “reading wars”. The heart of the conflict was a debate about the validity of what were termed “skills” vs. “literacy” approaches, which, amazingly, were seen as competing alternatives. 7 The scientists were seen as focused on “skills” (e.g., learning to read words and sentences accurately and fluently; vocabulary development), whereas educators emphasized developing “literacy” (the child's appreciation the various types and uses of written language, by individuals with diverse backgrounds, values and cultural traditions). Classroom time is a zero-sum game and so choices between skills and literacy had to be made. Moreover, teaching basic skills to beginning readers was thought to be counterproductive because it stifles children's natural curiosity about reading and their motivation to learn. This basic skills stuff may be necessary but it is also poisonous in large doses, so the child should be exposed to as little of it as possible. The traditional goal of teaching children to read has been replaced by coaching: encouraging the appreciation of and engagement in “multiple literacies.” Educational theorizing has gone “meta” about reading: there's little about how reading works (i.e., its neurocognitive bases), and much about how reading is used (various “literacy practices”) and by whom (various cultural/ethnic/language groups).

This conflict—which I am by no means overstating—arises from a failure to assume a genuinely developmental perspective. The act of reading and comprehending text involves the coordination of cognitive, linguistic, perceptual, motoric, memory and learning capacities. Understanding these capacities, how they develop, and how they are recruited in support of reading is obviously relevant to being able to help children become successful readers. What is relevant to teach (or “facilitate”) depends on where the child is on an extended developmental trajectory. The ability to read and comprehend words and their components is a basic, foundational skill. Helping children achieve this skill, without creating disinterest in reading, is the educational challenge. Acquisition of this foundation allows the child to benefit from other activities that promote further advancement: extended practice reading a variety of texts, with close checks on comprehension; reading texts for different purposes; gaining background knowledge relevant to what is being read. Socio-economic and cultural factors are highly relevant to the child's ability to benefit from schooling, but they do not change the nature of the reading process, or the kinds of knowledge and skills that need to be acquired.

3. Scientific literacy

The gap between the cultures ensures that people coming from the education side have little opportunity to gain an understanding of how research is conducted in relevant disciplines such as cognition, development, and neuroscience. Schools of education socialize prospective teachers into an ideology about children, learning, and reading. Prospective teachers are not exposed to other research that is relevant to their jobs, which is especially damaging given how difficult those jobs are. Educators are unprepared to engage this science in a serious way because they lack the tools to understand what is studied, how it is studied, what is found, what it means, and its relation to other kinds of research. This also leaves educators vulnerable to claims that are intuitively appealing but unproved, overhyped, or discredited. Educators embrace the importance of “critical thinking skills” and “background knowledge” in reading and learning, and so it is ironic when they are missing from discussions of research on reading and learning. I think that this deep ambivalence about the relevance of science to the educational mission explains seemingly contradictory features of educational culture such as the cherry-picking of selected findings, while at the same time discounting the relevance of basic research (e.g., Duke & Martin, 2011 ). I think it also explains why the single most influential educational theorist in America is Lev Vygotsky, who lived in the Soviet Union, wrote in Russian, died in 1934, and never saw an American classroom, or a television, computer, calculator, videogame or smartphone, yet educators are also looking to the latest findings from neuroscience for help (e.g., Willis, 2007 ). It is hard to know what Vygotsky, who founded the socio-cultural framework for education as an alternative to approaches based on psychology and biology, would have thought of this latest development.

Does It Matter?

The people who teach teachers and create curricula don't pay much attention to the science of reading, but is there reason to think that closer alignment of science and education would result in better outcomes? There have always been competing views about how reading should be taught, or, indeed, if it needs to be taught at all. People who have had vastly different educational experiences manage to become skilled readers. We know that teacher quality has a huge impact on educational outcomes (e.g., Hanushek & Rivkin, 2006 ) , but what about different ways of teaching reading?

It should matter. Reading is a learned skill, an “unnatural act” in Gough's memorable phrase. Some children find it easy to learn to read regardless of what happens in the classroom; many are well on their way by the onset of formal schooling. Other children will have difficulty learning to read regardless of what happens in the classroom because they are dyslexic: they have a developmental disorder that interferes with learning to read. Few teacher education programs provide any serious training related to developmental disorders such as dyslexia, how children at risk can be identified, and how such children can be helped. Whereas researchers are closing in on the neural and genetic bases of dyslexia ( Gabrieli, 2009 ), educational theorists are still debating whether dyslexia exists, and if it does, whether knowing that a child has the disorder should have any impact on classroom practices ( Elliott & Gibbs, 2008 ). Many of those children and adults who score poorly on national assessments are undoubtedly dyslexics whose condition has not been identified or addressed.

Between these extremes there is the great majority of children for whom how reading is taught matters a great deal. They are why we should care about what teachers are taught about reading. The main problem is that many of the basic assumptions about how children learn to read that have guided teacher education, classroom practices, and curriculum development have been contradicted by the basic research that lead to the Modern Synthesis. Beliefs about reading, learning, and development, reinforced over many years within the insular culture of schools of education, do not coincide with facts about reading, learning, and development uncovered using a variety of methods in laboratory and naturalistic settings. Rather than repeat details reviewed in sources I've already mentioned, let me try to capture the essence of the problem.

Everyone agrees that children have to acquire basic skills related to processing the visual code (letter recognition, learning about orthographic structure and the relationships between orthography and phonology, etc.), which provide a foundation for developing the ability to comprehend different kinds of texts for different purposes. Beyond this basic observation, there are two contradictory views.

Educators have assumed that basic skills are relatively easy to acquire, but comprehension is hard. Acquiring basic skills is mostly a matter of providing a literacy-rich environment with activities that engage and motivate the child. Learning to read was assumed to be like learning a spoken language. Children do not need to be explicitly instructed in how to read any more than they needed instruction in how to speak a first language. In practice—a Whole Language K-3 classroom—this meant de-emphasizing instruction related to acquiring basic skills. In the appropriate environment, full of “authentic” literature (rather than books written for the purpose of teaching reading), literacy activities focused on extended, “multisensory” engagement with a book (e.g., reading a book to the child, small groups of children reading the book aloud together, making a personal copy of the book, drawing pictures of the book, coloring the book, “writing” about the book using invented spelling, talking about the book, etc.), the child would discover the mechanics. Following John Dewey, discovering how reading works is assumed to have more value than being taught to read. The teacher's role is to promote literacy, not teach reading.

Comprehension, in contrast, was thought to be hard. The great fear was that children might develop basic skills and yet fail to comprehend texts. (Indeed it was thought that an initial focus on phonics would make it harder to become a good comprehender.) And so, inspired by theorists such as Frank Smith (1971 , now in its sixth edition), curricula focused on developing the child's explicit knowledge about text structure, types of inferences, the varied relationships between author and reader, the varied goals of reading, how to monitor comprehension and repair errors, and so on.

On the science side, the story is the exact opposite. Basic skills are difficult to acquire (mainly because of the partial and abstract way that writing systems represent spoken language) and thus the area where instruction matters most. Comprehension, in contrast, depends on extended experience using spoken and written language for varied purposes. Environments and activities that provide such experience can therefore promote comprehension skill. Ironically, this aspect of becoming a skilled reader more closely resembles spoken language acquisition than does the acquisition of basic skills.

Reading comprehension initially depends on the child's knowledge of spoken language. Learning a first language involves encoding detailed information about the statistical structure of the utterances to which the child is exposed ( Seidenberg, 1997 ). This information is continually updated through the lifespan, via language use ( Haskell, Thornton, & MacDonald, 2010 ). Comprehension abilities vary among children because there are huge individual differences with respect to the sheer amount of spoken language to which the child is exposed, and the range of vocabulary items and sentence structures it includes. Thus, reading comprehension can be improved by enriching the child's knowledge of spoken language ( Hoff, 2013 ).

Children who are given the instruction and support to acquire basic skills can advance to reading varied texts for varied purposes, learning from feedback about whether they are succeeding rather than explicit instruction in how to comprehend. Promoting skill development through engagement and feedback is different from teaching the child a meta-theory of comprehension. Eventually the relationship between spoken and written language becomes reciprocal and interactive. Knowledge of spoken language facilitates learning to read; the child can then acquire vocabulary and familiarity with diverse grammatical structures from reading as well as speech. What is learned from reading also contributes to the further development of spoken language skills. For skilled readers, the systems become closely knit, even at the neural level ( Pattamadilok et al., 2010 ).

In short, theorists on the education side had the instructional vs. experiential demands of acquiring basic skills vs. comprehension backward . Generations of teachers were then taught that the skills come naturally and that comprehension requires explicit instruction. That reversal has made learning to read harder for many children.

Finally, because of this two-cultures problem, there is very little opportunity to focus on how to best integrate basic research with educational practices. The science of reading is highly advanced but it does not come with prescriptions about educational practice attached. It is one thing to know about how children learn to read and another to translate that knowledge into classrooms. The NRP report again offers a good illustration of the zeitgeist. The report did a good job describing the main elements involved in learning to read, and the supporting evidence. It was not within the panel's mandate to address the educational issue, how these components could be affectively addressed within an integrated, multi-year reading curriculum. Thus the report described the importance of elements such as phonemic awareness and vocabulary in early reading, but not the levels of competence that are developmentally appropriate or how to assess them, or the effectiveness of different instructional methods. This created an enormous loophole. It is very easy to design curricula that can be said to conform to the recommendations of the NRP, simply by touching on all of the components they listed, even if only for a day. There has not been a serious dialogue about the pedagogical implications of the science summarized in the report, one of the major factors contributing to the science's lack of impact.

The Impact of Language Variation in the Educational Context

There cannot be a serious discussion of literacy issues in the US without considering the “achievement gap.” The term refers to disparities in academic performance between groups of individuals. It is mainly used with reference to minority groups – African Americans, Hispanics, Native Americans – compared to whites, but there are many other “gaps.” There are huge “gaps” associated with income disparities ( Reardon, 2011 ), and there are gaps for other groups such as first generation children of immigrants to the US compared to later generation children. Such “gaps” are seen in reading, math, science and in other areas, and on a variety of indices, including grades, standardized test scores, the kind of classes students take, high school and college completion rates, and so on (Washington et al., in press). My focus is on the achievement gap in reading and, again, whether what we have learned from our research could be brought to bear on it. 8

I am also going to limit attention to the gap for African Americans, even though they exist for groups defined in many other ways. Why single out this group? First, because this gap is a major issue for a very large number of affected individuals. Second, because it is part of a long history of racial inequality in the US. Third, because it has been the focus of attention from politicians, educators, and economists for many years ( Jencks & Phillips, 1998 ; Equity and Excellence Commission, 2013 ). Fourth, because this gap has been persistent. It has existed for as long as relevant data have been collected, with little change despite government efforts dating from the War on Poverty through No Child Left Behind to Race to the Top. Fifth, it is an area in which I am conducting research (Washington et al., in press). Finally, the various achievement gaps in our society have varying causes. Conditions and circumstances that are highly relevant to one group may be moot for another. Although this focus is warranted it is also essential to recognize it as part of a much broader phenomenon affecting many diverse groups of people.

The causes of the achievement gap in reading for African Americans (and other groups) are obviously complex and cannot be covered fully here (see, e.g., Richardson, 2008 ; Magnuson & Duncan, 2006 ; Barton & Coley, 2009 ; Washington et al., in press). The topic is also a sensitive one, having to do with generalizations about groups, within which there is enormous variation. What is said here cannot be assumed to apply to all members of a group, or to any individual within the group. It can also be harmful to raise issues about group differences in contexts that do not permit serious exploration and exchange of ideas. 9 I also know from experience that anything that is said about this issue, however well-reasoned, backed by evidence, and carefully stated with necessary qualifications attached can be spun for political purposes that researchers cannot control. My goal here is limited: to establish the relevance of research on reading and language to understanding and potentially ameliorating this gap. My only personal agenda is to encourage others to conduct research in the area, for the same reasons I became involved: because the issues are scientifically interesting; because existing research on reading and language acquisition is relevant; because there is a research gap insofar as the factors and conditions specific to African American children's reading acquisition are understudied; because this gap could be addressed by researchers who study other aspects of language acquisition and reading; and because the consequences of reading failures are so devastating.

Econometric Analyses: What Is Missing from this Picture?

The basis of the Black-White achievement gap in reading has been extensively studied by econometricians and sociologists. Several important analyses have utilized a large publicly-available data set, the Early Childhood Longitudinal Study ( http://nces.ed.gov/ecls ). These data were derived from extensive interviews with large numbers of individuals (about 20,000), supplemented with scores on assessment instruments. Consistent with other research, the ECLS data show that there is an achievement gap at the start of schooling: African American children are behind on measures of reading and pre-reading skills in kindergarten ( Duncan & Magnuson, 2005 ). Researchers have attempted to identify the bases of this gap by determining which factors in the data set account for the difference. Fryer and Levitt (2004) showed that six factors (out of over a hundred that were considered) accounted for the difference at the onset of schooling in this statistical sense: a composite measure of socioeconomic status, child's age at the start of kindergarten, birth weight, age of mother at time of first birth, whether the mother was a WIC (welfare) participant, and number of children's books in the home. These results are correlational, of course, and open to varied interpretation. This particular set of factors seems to be mainly tapping into SES and sequelae such as poorer health and health services, and fewer resources such as books in the home.

The results from additional waves of data collection yielded the surprising finding that the size of the gap increased through grade 3 ( Fryer & Levitt, 2006 ; see also Magnuson & Duncan, 2006 ). The causal interpretation is again unclear; schooling could either be exacerbating the gap, or the positive effects of schooling might be outpaced by the increasing impact of other factors. In either case, schooling was not acting as the “great equalizer”. 10 Moreover, the six factors that had accounted for the gap in kindergarten did not account for the gap in third grade, nor did any other factors in the data set. Fryer and Levitt wrote, “None of the explanations we examine[d], including systematic differences in school quality across races, convincingly explain the divergent trajectory of Black students.”

The fact that SES-related factors did not account for the increase in the gap is consistent with other findings indicating that the reading gap is not limited to lower income individuals. As noted above, performance on the NAEP is affected by SES, as indexed by the eligibility for subsidized lunch proxy. However, the black-white gap is highly consistent across the three levels of this measure (see Vanneman et al., 2009 , p. 33). Other studies have found that the gap exists for middle income blacks as well as low income ( Gosa & Alexander, 2007 ). Again it has to be emphasized that with overlapping distributions and imperfect correlations, there will be individuals who differ from these overall trends, including lower SES blacks whose reading achievement is on par with high-achieving, higher SES whites, a point that Magnuson and Duncun (2006 , p. 368) emphasized, noting that nearly a quarter of the black kindergarteners in the ECLS-K outscored the median for white students However, the group differences merit attention.

Here, then, is a puzzle. We are looking for a missing factor (or factors) with the following characteristics:

  • It contributes to the increasing deficit from K-3.
  • It affects individuals at different SES levels.
  • It is not captured by the measures included in data sets like the ECLS.

What is it?

One possibility is: language. There are two elements to consider. One is knowledge of spoken language, which varies across children, like many other skills and capacities. The other is the nature of the linguistic codes to which children are exposed. I will focus on different dialects of English, although similar issues arise regarding exposure to different languages. I will consider each of these elements—language-general and dialect-related—in turn.

There is very little data about children's language in large-scale data sets such as the ECLS, the NICHD Study of Early Child Care and Youth Development or the Children of the National Longitudinal Survey of Youth 79 survey. 11 The ECLS-K survey studied by Fryer and Levitt includes an item about whether English is spoken in the home and a “cognitive assessment” consisting of items taken from a variety of standardized tests, but no direct assessments of the characteristics of a child's language and linguistic environment. Yet there is a substantial body of evidence about the impact of these factors on children's school achievement, particularly reading ( McCardle et al., 2001 ). Children vary considerably with respect to knowledge of components of spoken language, including vocabulary size, morphology, and syntax ( Bates et al., 1995 ). People often refer to differences in verbal “ability,” but the relative richness of children's language is affected by exogenous factors and so this term seems unsuitable. Some refer to spoken language “quality”, but in this context the term evokes the discredited idea that African American English is inherently inferior. Lacking a better term I will refer wherever possible to differences in general knowledge of spoken language. These differences could arise from constitutional, environmental, and socio-cultural factors.

The most widely studied such factor is SES. Differences in linguistic input associated with SES were documented in Risley and Hart's famous study (1995; see also Hansen & Joshi, 2010 ; Hoff, 2013 ). Children will have difficulty learning words, grammatical structures, and discourse conventions to which they are not exposed. Vocabulary in particular is strongly related to progress in learning to read ( NRP, 2000 ). In short, children's success in making the transition into reading depends heavily on their knowledge of spoken language, which varies across individuals and is associated with differences in SES ( Fernald & Marchman, 2011 ). African American children are overrepresented at the lower end of the SES distribution; therefore, they will be disproportionately subject to the effects of low SES on language.

This argument is inconclusive, however. The SES-related factors that Fryer and Levitt identified may be relevant, in part, because of their association with differences in language input, but this cannot be determined because the ECLS-K data set does not include measures of child or caregiver vocabulary (or other aspects of spoken language). The relations between their six factors and the child's knowledge of spoken language are indirect at best. Moreover, these factors accounted for differences at the onset of schooling, but not the growth in the gap through grade 3. Thus, whatever aspects of the linguistic environment they might be picking up are not sufficient to explain the increase. Finally, there is the fact that the achievement gap is not limited to the lower SES cohort. Disentangling the complex relationship between SES and the achievement gap continues to be the focus of research ( Duncan & Magnuson, 2012 ). A cautious reading of the existing literature suggests that there must be other factors involved and that the contributions of language variability need to be assessed more directly.

A further consideration not addressed by studies such as the ECLS is the nature of the linguistic code to which the child is exposed, in particular dialect. Dialects are variants of a language, spoken by individuals grouped by region, ethnicity, race, income, and other factors (Chambers & Trudgill, 1998). Every native speaker of a language learns a dialect of that language. English has many dialects, identifiable at different grain sizes (e.g., American vs. British English; regional dialects in these countries, etc.). In the US, one major division is between so-called “standard” or “mainstream” American English (SAE) and African American English (AAE). As in other countries, which dialect is treated as “standard” is not a linguistic issue but rather is determined by demographic, cultural and political considerations. AAE and SAE overlap—they are both versions of English—but also differ with respect to specific elements of phonology, morphology, lexicon, syntax, and discourse/pragmatics ( Rickford, 1999 ). Although like others I will refer to AAE, it is important to recognize that it has regional variants, as does SAE ( Green, 2002 ; Wolfram & Schilling-Estes, 2006). Moreover, speakers vary in the extent to which they use characteristic features of AAE (a dimension termed dialect density; Thompson et al., 2004 ); thus the extent to which AAE differs from SAE also varies. AAE is used by most African Americans, at varying densities, across SES levels ( Washington, 1996 ). The question then is whether use of AAE contributes to the reading gap.

Research on this question extends back many decades (see Washington et al., in press, for more detailed review). One issue is whether an AAE speaker would be disadvantaged because the dialect is deficient in some manner. This issue was decisively resolved by the basic linguistic research on AAE conducted by Labov (1972) and subsequently many others ( Rickford et al., 2004 ). This research appropriately situated AAE in the context of dialectal variation as it occurs in languages around the world. One of the great achievements of this research was to establish how unexceptional AAE is as an example of linguistic variation (Chambers & Trudgill, 1998). Whether dialect use affects school achievement is not a question about the linguistic status of AAE. The unresolved question is whether use of the minority dialect has an impact because of the conditions surrounding its use. Specifically, AAE usage could affect the child's ability to benefit from educational experiences because of sociological and cultural factors (e.g., AAE is a “low status” dialect; books are written in SAE; acquiring skill in SAE is an educational goal in American schools). It is because of these conditions that differences between the dialects are relevant, not because they are linguistically significant.

Labov's research stimulated considerable research on the possible impact of AAE on the development of African American children's reading skills. Most of the early studies found that dialect usage had little effect on comprehension (Washington et al., in press). Thus it was concluded that what mattered was the quality of linguistic experience, independent of dialect.

These studies were thought to have put the issue of AAE's impact on reading acquisition to rest, but they did so prematurely in my view. The early studies do not hold up well by modern research standards and the strong conclusions that were based on them need to be re-examined. In recent years, researchers have begun to reopen the issue, using what has been learned about reading, language, development, and cognition over the past several decades to generate more specific hypotheses that can be tested using more powerful research methods. The topic is still understudied and many empirical questions are unresolved.

One major question is this: if language variation is a major element in the achievement gap, to what extent does it involve language-general vs. dialect-related aspects of language? Early reading achievement is closely related to knowledge of spoken language. This relationship is general, applying across languages and, within a language, across dialectal variants. Whether young children differ with respect to skills such as phonological awareness, the ability to analyze spoken words at the phonemic level, vocabulary size, and spoken language comprehension should matter, not which variant of a language they happen to speak ( Terry, in press ; Terry & Scarborough, 2010).

It is also possible that there could be dialect-related effects on reading outcomes. Such effects could arise from a variety of factors that emerge when linguistic differences between dialects converge with extralinguistic factors related to the educational and socio-cultural conditions under which dialects are used. The language-general and dialect-related possibilities are not mutually exclusive. They also may not be independent: for example, the acquisition of a language-general skill could be affected by dialect-related factors. General and dialect-related influences may also be difficult to differentiate because a child's general language ability is manifested in the use of a particular dialect. Finally, there may be advantages to exposure to multiple dialects, analogous to those associated with exposure to different languages ( Bialystok et al., 2009 ), although this possibility is rarely considered.

These issues can be illustrated with respect to vocabulary knowledge, a factor that has been the focus of much research and is known to have a major impact on reading acquisition. It is a fact about languages that they consist of inventories of words (among other elements). Vocabulary size, however, is a characteristic of a child—what he/she knows about this component of spoken language—not the dialect that is spoken. This observation suggests that research should focus on the child's knowledge of properties of spoken language, such as vocabulary, irrespective of dialect.

A potential complicating factor is that differences in language background—e.g., use of a minority dialect, varying exposure to and knowledge of the mainstream dialect, dialect usage in the school context—could affect the child's acquisition of “general” elements of spoken language. A factor such as vocabulary size needs to be considered not just as a quantitative predictor of reading outcomes, but also with respect to a theory of how this knowledge is acquired. It could then be determined whether or to what extent specific aspects of dialect experience matter. Is vocabulary acquisition affected by properties of a dialect, or the need to accommodate two dialects? Are such effects positive or negative or both? Do the effects differ depending on the child's point in development? Are they modulated by individual differences in cognitive capacities such as executive function? These issues are not well understood. The bilingual literature suggests they are worth addressing, however. Many studies have shown that preschool-aged bilingual children have smaller vocabularies in each language than comparable monolingual speakers ( Bialystok et al., 2005 ), which then has an impact on learning to read in one of the languages. These effects also arise from conditions relevant to children's learning rather than properties of the languages. They also occur across SES levels. The bilingual burden is by no means insuperable, but the developmental time course may be affected, creating another “gap”, but also the emergence of bilingual advantages ( Bialystok et al., 2009 ). Analogous issues may arise for AAE speaking children who are learning to read in SAE.

As an illustration, consider the optional deletion of final consonants in pronouncing some words in AAE. One consequence is that a given word (such as COLD) can be pronounced differently, at the phonemic level, in the two dialects. Deletion of final consonants also creates additional homophony: words such as COLD and COAL have different pronunciations (at the phonemic level) in SAE but they can be homophonic in AAE. Do these differences between the dialects have any impact on language learning or reading? Or are the differences inconsequential? I don't think we know. The existence of alternative pronunciations across dialects could create a more complex word learning problem, or it might be no more difficult than assimilating differences in pronunciation arising from pitch, speech rate, and so on. An increase in the number of homophones might facilitate vocabulary acquisition (fewer distinct phonological word forms to learn) or make it harder (because the child has to use other mechanisms to disambiguate homophones). Or the functional impact could be trivial. These unanswered questions suggest that it would be premature to treat vocabulary size as a general linguistic factor independent of dialect experience.

Given the limited evidence that is available, I believe that it would be a recapitulation of an earlier mistake to conclude at this point that dialect experience has no significant impact on reading or other aspects of school achievement. As Snow et al. (1991) noted, differences between home and school language could affect children's learning. For speakers of the mainstream dialect, the home and school dialects are the same. For speakers of the minority dialect AAE, the home and school dialects differ in varying degrees. Thus, dialect use is consistent across contexts for one group but not the other. AAE speakers have to learn about the mainstream dialect and use both dialects at the same time they are learning to read, write, and do arithmetic. SAE speakers do not have the additional language-related demands. Speakers of the minority dialect clearly have to do more work in order to succeed. They are nonetheless assessed against the same achievement milestones as SAE speakers. If this analysis is correct, it means that the achievement gap has been intractable because it is built in, guaranteed by prevailing circumstances unrelated to the linguistic validity of the dialect or the capacities of the child. 12 Stated another way, if by analogy to the Fryer and Levitt analysis the overhead associated with accommodating dialect differences were somehow factored out of the natural experiment, the gap in the first years of schooling would greatly narrow rather than grow.

Differences between dialects could potentially affect children's ability to benefit from classroom experience in numerous ways. For example, an SAE-speaking teacher will pronounce many words differently from the child and use different morphological and syntactic constructions. The additional processing and attentional demands associated with comprehending utterances in the less familiar dialect and switching between dialects could be expected to interfere with the child's opportunity to learn from what is being said. The impact would be exacerbated by the fact that the classroom context is a noisy one (literally, and in the information theoretic sense), degrading the quality of linguistic signals. The effect would be similar to imposing a delay on the on-line processing of spoken utterances. These demands could at the same time promote the development of other capacities such as executive function. There are opportunities for communication failures from the opposite side as well: a teacher who speaks SAE and does not know AAE may have more difficulty understanding the AAE speaking child. The teacher may also use discourse conventions (e.g., indirect speech acts) with which the child is unfamiliar, or misinterpret the child's own discourse conventions.

Then there are specific ways in which dialect differences could affect learning to read. The child speaks one dialect but books and other materials are written in the other dialect, again imposing additional learning and processing demands compared to the child for whom the same dialect is used for both. A useful comparison is to hearing-impaired children who are fluent, native signers of ASL, who have high verbal skills, with large vocabularies and language that exhibits a high degree of syntactic variety and complexity—who are nonetheless poor readers. There is an “achievement gap” in reading for the hearing-impaired, due in part to the fact that they are learning to read in English, a different language. There are important differences between these two situations (e.g., the ASL-English differences are greater than the differences between English dialects; the use of different dialects is not associated with presence or absence of a perceptual deficit such as hearing loss), but the analogy is apposite.

And what about acquiring basic reading skills? The beginning reader's initial challenge is to learn how the spoken language they know relates to the written code they are learning. Making this connection is difficult for many children, for reasons that have been investigated in great detail ( NRP, 2000 ). Reading an alphabet involves learning to treat spoken words as if they consist of discrete phonemes. Units in the written code (letters and digraphs) can then be mapped onto units in the spoken code (phonemes). Making this abstraction is difficult for many children. The task is further complicated by inconsistencies in the mappings between spelling and sound in English. The task becomes even more complex when a substantial number of words are pronounced differently at the phonemic level in the two dialects. Consider just the subset of words in AAE in which the final consonant can be dropped (e.g., GOLD→/goυl /, BEST→/bes/). A teacher explains that the word “gold” is spelled G-O-L-D. For an SAE speaking child this is a lesson about the alphabetic principle and the correspondences between four letters and four phonemes. What is being taught an AAE speaking child who pronounces the word /goυl/? That there are different ways to pronounce the word? That if the spelling maps onto one pronunciation, the final letter is pronounced /d/, whereas for the other it is silent? The alternative pronunciations create additional inconsistencies in the mappings between spelling and sound. The learning problem is further complicated by the fact that this deletion is not obligatory, and thus may be used for a given word only some of the time, or for only some words in a similarly spelled neighborhood (e.g., -OLD words).

In short, the need to accommodate both dialects may place a variety of additional burdens on young learners. The potential for dialect-related factors to affect learning—and the need to determine where differences between the dialects do and do not have a significant impact, positive or negative—does not invalidate the relevance of dialect-independent variability in spoken language skills. Plainly, both could exert influence, creating in the worst case a debilitating double-whammy. However, the extent to which such effects occur and how much impact they have are not well understood.

What is striking about dialect is how poorly it is addressed in America compared to other countries. Dialect variation is not specific to African Americans or English, but dialect differences seem to have greater prominence in this country than elsewhere. There are major dialect variants in countries such as Finland and Germany where literacy levels are higher than in the US. Although each situation is different in detail, the main challenges are the same. It appears that these countries do a better job of acknowledging and accommodating dialect differences. Are educators more familiar with dialect issues and their potential impact? Does teacher training include dialect-related issues? Is there less dialect-related prejudice in these countries? Is there greater exposure to alternative dialects prior to the onset of schooling? We know that what works in one country cannot simply be ported to another where relevant circumstances are different. Nonetheless, there is information to be gained from examining how such issues play out in other countries and languages. This effort might suggest ways of changing the culturally-determined conditions that contribute to the achievement gap.

Looking to the future, research on the achievement gap could take a page from research on reading and language disorders. Research on dyslexia, for example, has shifted from a focus on single causes (e.g., a visual or phonological deficit) to the view that outcomes arise from the aggregate effects of a set of risk factors ( Snowling & Hulme, 2011 ; Morris et al., 2010). Each factor is probabilistic in the sense that it does not itself guarantee a particular outcome. The factors also vary in degree of severity and relative impact. Together these factors yield a range of behavioral outcomes. The crucial linkage between the risk factors and outcomes is provided by theories of reading that specify major subskills and how they are learned. By analogy, the poor reading outcomes seen in the “achievement gap” arise from a variety of risk factors that are also probabilistic, vary in severity and impact, interact in complex ways, are differentially amenable to intervention, and yield a broad range of individual outcomes. Risk factors relevant to African American children in American schools include knowledge of spoken language (“general” language skills), language experience (e.g., dialect usage, exposure to and knowledge of the alternative dialect, the cognitive demands of dialect switching), adequacy of educational responses to language variation, and poverty, among others. There is a further need to link these risk factors to specific components of reading and how they are acquired.

Conclusions

Reading failures arise from multiple causes. My goal has been to suggest that this serious societal issue can further benefit from the kinds of research that we conduct as scientists who study reading and language. Reading is often viewed as secondary to spoken language and of very little linguistic interest. Reading did not evolve in the species, it is true, but once the technology became available to many people, it became as central an expression of our capacity for language as speech, greatly changing the ways that language can be used. There are scientists who study reading qua reading, and we have made considerable progress in understanding this skill. But questions about how reading works and the determinants of reading skill invariably come back to issues about spoken language. I've tried to suggest that our basic research is highly relevant to a societal problem of enormous importance. The challenges are daunting, and the need is great.

Acknowledgments

I am greatly indebted to Julie Washington, who stimulated my interest in dialect issues and the achievement gap, took me to school, and gave me the opportunity to become involved in research on the topic. She should not be held responsible for the contents of the article and especially any mistakes, but she has greatly influenced my thinking. I also thank Katherine Magnuson for guidance in navigating the socio-economic literature; she also provided very helpful feedback on a previous version, as did Heidi Feldman, Richard Aslin, and Maryellen MacDonald. I am grateful to Jon Willits for collecting the data in Figure 1 . Finally I thank the journal editor, Cindy Fisher, for her patience in dealing with a manuscript that was always almost finished. Preparation of this article was supported by NICHD grant R24D075454 (Julie Washington, Nicole Patton Terry, and Mark Seidenberg, PIs).

1 See the PISA summary document available here: http://www.oecd.org/pisa/pisaproducts/pisa2009/pisa2009keyfindings.htm .

2 See, for example, the summary report for the 2011 reading assessment, available here: http://nces.ed.gov/nationsreportcard/pubs/main2011/2012457.asp

3 Basic facts about Finnish elementary education are available here: http://www.oecd.org/pisa/pisaproducts/46581035.pdf .

4 Is Hebrew the outlier writing system? Seemingly contrary to my analysis, it is morphologically complex but also orthographically deep in the default, unpointed form. Note, however that children learn to read using the shallow form in which vowels are indicated by diacritics ( niqqud ). Learning to read using the unpointed form would be vastly more difficult (though perhaps it was achieved by the ancient scribal elite prior to the development of the diacritic system).

5 The Mann quote (which I first encountered in Adams, 1990 ) is from an 1844 report he prepared as Secretary (head) of the Massachusetts Board of Education in which he was highly critical of the local schools, comparing them unfavorably to the classrooms he had observed in Prussia and Scotland (shades our modern-day envy of educational practices in Finland and Shanghai!). Greatly offended, schoolmasters from the Boston public schools published a rejoinder in which they remarked that “Our dissent from [Mann's] views arises from an honest conviction that, if adopted, they would retard the progress of sound learning.” Mann was advocating what later became known as the whole-word or “look and say” method, which involves memorizing words as patterns, without regard to the functions of the component letters. The Boston educators favored a “phonetic” teaching method. Their take-down of Mann's “new method” was thorough and incisive but settled nothing. The arguments on both sides will be easily recognizable to anyone familiar with the “Reading Wars” of the past 30 years. All the documents (the sides went back and forth a few times) are available as Ebooks on Google Play and highly recommended.

6 See Allington and Woodside-Jiron (1999) , who believe that many of the research findings that contradict their own views were the product of research funded by Reid Lyon, an official at NICHD, as part of an anti-education political agenda. The founding document for this political movement, they claim, is Grossen (1997) , an obscure 22 page review of 30 years of reading research funded by NICHD. Allington and Woodside-Jiron's paranoia is so keenly focused on NICHD that they ignore the mass of similar findings from research conducted in many other countries. The same conclusions about learning to read are found in both American reports such as the NRP (2000) and the British Rose Report ( Rose, 2006 ). It would be easier to dismiss Allington's campaign against reading science (see also Allington, 2002 ) were he not a leading figure in reading education, former president of the International Reading Association, former president of the National Reading Conference, and a member of the “Reading Hall of Fame”, http://www.readinghalloffame.org ).

7 In the current climate, everyone has to favor a “balanced” approach to reading instruction, acknowledging the importance of both skills and literacy. Having seen and comprehended the writing on the wall, organizations that had gone to the mat in support of “literacy” approaches such as Whole Language have turned out guidelines for “balanced literacy” instruction (see, e.g., Cowen, 2003 , for an example, and Moats, 2007 , for a critique of such efforts).

8 Basic data on achievement gaps in reading, as measured on the National Assessment of Academic Progress (NAEP), can be found in the Executive Summary of the 2011 results, pp. 15 and 44 ( http://nces.ed.gov/nationsreportcard/pubs/main2011/2012457.asp ).

9 As exemplified by then-Harvard President Lawrence Summers' conjectures about possible sex-linked genetic differences in mathematical aptitude ( http://www.harvard.edu/president/speeches/summers_2005/nber.php ). I believe that an article in a journal such as this one provides an appropriate context.

10 These data give the lie to a cherished belief. Then: “Education then, beyond all other devices of human origin, is a great equalizer of the conditions of men — the balance wheel of the social machinery.” Horace Mann, 1848 . And now: “In America, education is still the great equalizer.” Arne Duncan, U.S. Secretary of Education, 2011 ( http://www.ed.gov/news/speeches/remarks-us-secretary-education-arne-duncan-he-education-trust-conference ). We should also be considering whether education, as it occurs in American schools and as it is funded, exacerbates differences between groups.

11 NICHD Study of Early Child Care and Youth Development, http://www.nichd.nih.gov/research/supported/Pages/seccyd.aspx ; Children of the National Longitudinal Survey of Youth 79 survey, http://www.bls.gov/nls/nlsy79.htm .

12 I owe this observation to Julie Washington, who made the point with great clarity and impact. The finish line may be in the same location, but paths to getting there are not of equal length or difficulty.

  • Adams M. Beginning to read. Cambridge, MA: MIT Press; 1990. [ Google Scholar ]
  • Allington RL. Big brother and the national reading curriculum: how ideology trumped evidence. Portsmouth, NH: Heinemann; 2002. [ Google Scholar ]
  • Allington RL, Woodside-Jiron H. The politics of literacy teaching: How “research” shaped educational policy. Educational Researcher. 1999; 28 :4–13. [ Google Scholar ]
  • Aro M, Wimmer H. Learning to read: English compared to six more regular orthographies. Applied Psycholinguistics. 2003; 24 :621–635. [ Google Scholar ]
  • Association Of Masters Of The Boston Public Schools. Remarks on the seventh annual report of the Hon Horace Mann, secretary of the Massachusetts Board of Education. Boston, MA: Charles C. Little and James Brown; 1844. [ Google Scholar ]
  • Au KH. Social constructivism and the school literacy learning of students of diverse backgrounds. Journal of Literacy Research. 1998; 30 :297–319. [ Google Scholar ]
  • Barton PE, Coley RJ. Parsing the achievement gap II. Policy Information Center, Educational Testing Service 2009 [ Google Scholar ]
  • Bates E, Dale PS, Thal D. Individual differences and their implications for theories of language development. In: Fletcher P, MacWhinney B, editors. Handbook of child language. Oxford, UK: Basil Blackwell; 1995. pp. 96–151. [ Google Scholar ]
  • Bialystok E, Craik FI, Green DW, Gollan TH. Bilingual minds. Psychological Science in the Public Interest. 2009; 10 :89–129. [ PubMed ] [ Google Scholar ]
  • Bialystok E, Luk G, Kwan E. Bilingualism, biliteracy, and learning to read: Interactions among languages and writing systems. Scientific studies of reading. 2005; 9 :43–61. [ Google Scholar ]
  • Boring EG. A history of introspection. Psychological Bulletin. 1953; 50 :169–189. [ PubMed ] [ Google Scholar ]
  • Charity AH, Scarborough HS, Griffin DM. Familiarity with school English in African American children and its relation to early reading achievement. Child Development. 2004; 75 :1340–1356. [ PubMed ] [ Google Scholar ]
  • Coles G. Misreading Reading: The Bad Science That Hurts Children. Portsmouth, NJ: Heinemann; 2000. [ Google Scholar ]
  • Cowen JE. A Balanced Approach to Beginning Reading Instruction: A Synthesis of Six Major U S Research Studies. International Reading Association; 2003. [ Google Scholar ]
  • Cox D, Meyers E, Sinha P. Contextually evoked object-specific responses in human visual cortex. Science. 2004; 304 :115–117. [ PubMed ] [ Google Scholar ]
  • Cremin LA. American education: the metropolitan experience, 1876-1980. Harper Collins; 1988. [ Google Scholar ]
  • Daniels PT, Bright W. The world's writing systems. Oxford University Press; 1996. [ Google Scholar ]
  • Duke NK, Martin NM. 10 things every literacy educator should know about research. The Reading Teacher. 2011; 65 :9–22. [ Google Scholar ]
  • Duncan GJ, Magnuson KA. Can family socioeconomic resources account for racial and ethnic test score gaps? The future of children. 2005; 15 :35–54. [ PubMed ] [ Google Scholar ]
  • Duncan GJ, Magnuson K. Socioeconomic status and cognitive functioning: moving from correlation to causation. Wiley Interdisciplinary Reviews: Cognitive Science. 2012; 3 :377–386. [ PubMed ] [ Google Scholar ]
  • Durgunoğlu AY. How the language's characteristics influence Turkish literacy development. In: Joshi M, Aaron PG, editors. Handbook of orthography and literacy. Mahwah, NJ: Lawrence Erlbaum Associates; 2006. pp. 219–230. [ Google Scholar ]
  • Elliott JG, Gibbs S. Does dyslexia exist? Journal of Philosophy of Education. 2008; 42 :475–491. [ Google Scholar ]
  • Equity and Excellence Commission. For each and every child—A strategy for equity and excellence. 2013 http://www.foreachandeverychild.org/60565_EEC_(508)_rev.pdf .
  • Fernald A, Marchman VA. Experience, Variation, and Generalization: Learning a First Language (Trends in Language Acquisition Research) Amsterdam: John Benjamins; 2011. Causes and consequences of variability in early language learning; pp. 181–202. [ Google Scholar ]
  • Finn CE. (2010). A Sputnik moment for U.S. Education. Wall St Journal. 2010 Dec 8; [ Google Scholar ]
  • Fryer RG, Levitt SD. Understanding the black-white test score gap in the first two years of school. Review of Economics and Statistics. 2004; 86 :447–464. [ Google Scholar ]
  • Fryer RG, Levitt SD. The black-white test score gap through third grade. American Law and Economics Review. 2006; 8 :249–281. [ Google Scholar ]
  • Gabrieli JDE. Dyslexia: A new synergy between education and cognitive neuroscience. Science. 2009; 325 :280–283. [ PubMed ] [ Google Scholar ]
  • Gee JP. Thinking, learning, and reading: The situated sociocultural mind. In: Kirshner DI, Whitson JA, editors. Situated cognition: Social, semiotic, and psychological perspectives. 1997. pp. 235–259. [ Google Scholar ]
  • Gibson E, Levin H. The psychology of reading. Cambridge, MA: MIT Press; 1978. [ Google Scholar ]
  • Goldin-Meadow S, Mayberry R. How Do Profoundly Deaf Children Learn to Read? Learning Disabilities Research & Practice. 2001; 16 :222–229. [ Google Scholar ]
  • Gosa TL, Alexander KL. Family Disadvantage and the Educational Prospects of Better Off African American Youth. Teachers College Record. 2007; 109 (2):285–321. [ Google Scholar ]
  • Gough PB, Hillinger ML. Learning to read: an unnatural act. Bulletin of the Orton Society. 1980; 30 :179–196. [ Google Scholar ]
  • Government Accounting Office. Poverty in America. 2007 Report GAO-07-344. [ Google Scholar ]
  • Green LJ. African American English: a linguistic introduction. Cambridge University Press; 2002. [ Google Scholar ]
  • Grossen B. 30 years of research: What we now know about how children learn to read. 1997 http://www.eric.ed.gov/PDFS/ED415492.pdf .
  • Hanley JR, Masterson J, Spencer L, Evans D. How long do the advantages of learning to read a transparent orthography last? An investigation of the reading skills and reading impairment of Welsh children at 10 years of age. Quarterly Journal of Experimental Psychology. 2004; 57 :1393–1410. [ PubMed ] [ Google Scholar ]
  • Hansen K, Jones E, Joshi H, Budge D. (2010) Millennium Cohort Study Fourth Survey: A User's Guide to Initial Findings. 2nd. London: Centre for Longitudinal Studies; Dec, 2010. [ Google Scholar ]
  • Hanushek EA, Rivkin SG. Teacher Quality. In: Hanushek E, Welch F, editors. Handbook of the Economics of Education. Elsevier; 2006. pp. 1051–1078. [ Google Scholar ]
  • Hart B, Risley TR. Meaningful differences in the everyday experience of young American children. Baltimore, MD: Brookes Publishing Company; 1995. [ Google Scholar ]
  • Haskell TR, Thornton R, MacDonald MC. Experience and grammatical agreement: Statistical learning shapes number agreement production. Cognition. 2010; 114 :151. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hoff E. Interpreting the early language trajectories of children from low-SES and language minority homes: implications for closing achievement gaps. Developmental Psychology. 2013; 49 :4–14. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Hoover WA, Gough PB. The simple view of reading. Reading and Writing. 1990; 2 :127–160. [ Google Scholar ]
  • Hoxhallari L, van Daal VHP, Ellis NC. Learning to read words in Albanian: A skill easily acquired. Scientific Studies of Reading. 2004; 8 :153–66. [ Google Scholar ]
  • Hulme C, Snowling M. Children's Reading Comprehension Difficulties: Nature, Causes, and Treatments. Current Directions in Psychological Science. 2011; 20 :139–142. [ Google Scholar ]
  • Jencks C, Phillips M. America's next achievement test: Closing the black-white test score gap. American Prospect. 1998; 9 :44–53. [ Google Scholar ]
  • Jorm AF, Share DL. Phonological recoding and reading acquisition. Applied Psycholinguistics. 1983; 4 :103–147. [ Google Scholar ]
  • Kahneman D. Thinking, fast and slow. Farrar, Straus and Giroux; 2011. [ Google Scholar ]
  • Katz L, Frost R. The reading process is different for different orthographies: The orthographic depth hypothesis. In: Frost R, Katz L, editors. Orthography, Phonology, Morphology, and Meaning. Amsterdam: Elsevier North Holland Press; 1992. pp. 67–84. [ Google Scholar ]
  • Krashen S. More smoke and mirrors: A critique of the National Reading Panel report on fluency. The Phi Delta Kappan. 2001; 83 :119–123. [ Google Scholar ]
  • Lonigan CJ, Shanahan T. Developing Early Literacy: Report of the National Early Literacy Panel Executive Summary A Scientific Synthesis of Early Literacy Development and Implications for Intervention. National Institute for Literacy; 2009. [ Google Scholar ]
  • Lesgold A, Welch-Ross M, editors. Improving adult literacy: Options for practice and research. National Research Council; 2012. [ Google Scholar ]
  • Liberman IY, Shankweiler D, Liberman AM, Fowler C, Fischer FW. Phonetic segmentation and recoding in the beginning reader. In: Reber AS, Scarborough DL, editors. Toward a psychology of reading: the proceedings of the CUNY conference. Hillsdale, NJ: Erlbaum; 1977. pp. 207–225. [ Google Scholar ]
  • Lindgren SD, deRenzi E, Richman LC. Cross-national comparisons of developmental dyslexia in Italy and the United States. Child Development. 1985; 56 :1404–1417. [ PubMed ] [ Google Scholar ]
  • Magnuson KA, Duncan GJ. The role of family socioeconomic resources in the black–white test score gap among young children. Developmental Review. 2006; 26 :365–399. [ Google Scholar ]
  • Mann H. Seventh annual report. Boston: Dulton & Wentworth; 1844. [ Google Scholar ]
  • Mann H. Twelfth annual report The republic and the school: Horace Mann on the education of free men. 1848:79–112. [ Google Scholar ]
  • McCardle P, Scarborough HS, Catts HW. Predicting, explaining, and preventing children's reading difficulties. Learning Disabilities Research & Practice. 2001; 16 :230–239. [ Google Scholar ]
  • Mirković J, MacDonald MC, Seidenberg MS. Where does gender come from? Evidence from a complex inflectional system. Language and Cognitive Processes. 2004; 20 :139–168. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Mirković J, Seidenberg MS, Joanisse MF. Rules vs. statistics: Insights from a highly inflected language. Cognitive Science. 2011; 35 :638–681. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Moats L. Whole language hijinks. Thomas B. Fordham Institute; 2007. http://www.edexcellence.net/publications/wholelanguage.html . [ Google Scholar ]
  • Moje EB, Luke A. Literacy and identity: Examining the metaphors in history and contemporary research. Reading Research Quarterly. 2009; 44 (4):415–437. [ Google Scholar ]
  • Morris RD, Lovett MW, Wolf M, Sevcik RA, Steinbach KA, Frijters JC, Shapiro MB. Multiple-component remediation for developmental reading disabilities: IQ, socioeconomic status, and race as factors in remedial outcome. Journal of Learning Disabilities. 2012; 45 :99–127. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Nation K, Cocksey J. The relationship between knowing a word and reading it aloud in children's word reading development. J Experimental Child Psychology. 2009; 103 :296–308. [ PubMed ] [ Google Scholar ]
  • National Assessment of Educational Progress. Reading 2011, Executive Summary. 2011 http://nces.ed.gov/nationsreportcard/pubs/main2011/2012457.asp .
  • National Reading Panel. Teaching children to read. Washington, D.C.: National Institute of Child Health and Human Development; 2000. http://www.nationalreadingpanel.org . [ Google Scholar ]
  • Pattamadilok C, Knierim IN, Duncan KJK, Devlin JT. How does learning to read affect speech perception? The Journal of Neuroscience. 2010; 30 :8435–8444. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Pugh KR, Landi Nicole, Preston Jonathan L, Mencl W Einar, Austin Alison C, Sibley Daragh, Fulbright Robert K, Seidenberg Mark S, Grigorenko Elena L, Constable R Todd, Molfese Peter, Frost Stephen J. The relationship between phonological and auditory processing and brain organization in beginning readers. Brain & Language. 2012 Available online 7 May 2012. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Pennington BF. From single to multiple deficit models of developmental disorders. Cognition. 2006; 101 :385–413. [ PubMed ] [ Google Scholar ]
  • Rasinski T, Yildirim K, Nageldinger J. Building fluency through the phrased text lesson. The Reading Teacher. 2011; 65 :252–255. [ Google Scholar ]
  • Ravitch D. Left behind: A century of battles over school reform. NY: Simon and Schuster; 2000. [ Google Scholar ]
  • Ravitch D. The death and life of the great American school system: How testing and choice are undermining education. Basic Books; 2011. [ Google Scholar ]
  • Reardon S. The widening academic achievement gap between the rich and the poor: New evidence and possible explanations. In: Duncan GJ, Murnane RJ, editors. Whither Opportunity? Rising Inequality, Schools, and Children's Life Chances. Russell Sage Foundation; 2011. pp. 91–116. [ Google Scholar ]
  • Richardson Elaine. African American literacies. In: Street B, Hornberger NH, editors. Encyclopedia of Language and Education. 2nd. Vol. 2. Springer; 2008. pp. 335–346. Literacy. [ Google Scholar ]
  • Rickford JR. African American vernacular English: Features, evolution, educational implications. Oxford, UK: Blackwell; 1999. [ Google Scholar ]
  • Rickford JR, Sweetland J, Rickford AE. African American English and other Vernaculars in Education: A Topic-Coded Bibliography. Journal of English Linguistics. 2004; 32 :230–320. [ Google Scholar ]
  • Rose J. Independent review of the teaching of early reading: Final report. London: Department for Education and Skills; 2006. Available at www.standards.dfes.gov.uk/phonics/rosereview . [ Google Scholar ]
  • Scribner S, Cole M. The psychology of literacy. Cambridge: Harvard University Press; 1981. [ Google Scholar ]
  • Seidenberg MS. Language acquisition and use: Learing and applying probabilistic constraints. Science. 1997; 275 :1599–1604. [ PubMed ] [ Google Scholar ]
  • Seidenberg MS. Reading in different writing systems: One architecture, multiple solutions. In: McCardle P, Ren J, Tzeng O, editors. Dyslexia Across languages: Orthography and the Gene-Brain-Behavior Link. Paul Brooke Publishing; 2011. [ Google Scholar ]
  • Seidenberg MS. What happened to reading. New York: Basic Books; forthcoming. [ Google Scholar ]
  • Seidenberg MS, McClelland JL. A distributed, developmental model of word recognition and naming. Psychological Review. 1989; 96 :523–568. [ PubMed ] [ Google Scholar ]
  • Smith Frank. Understanding reading: A psycholinguistic analysis of reading and learning to read. New York: Holt, Rinehart & Winston; 1971. [ Google Scholar ]
  • Snow C. Reading for Understanding: Toward an R & D Program in Reading Comprehension. RAND Corporation report; 2002. http://www.rand.org/content/dam/rand/pubs/monograph_reports/2005/MR1465.pdf . [ Google Scholar ]
  • Snow CE, Burns MS, Griffin P, editors. Preventing reading difficulties in young children. National Academies Press; 1998. [ Google Scholar ]
  • Snow C, Baines W, Chandler J, Goodman I, Hemphill L. Unfulfilled expectations: Home and school influences on literacy. Cambridge, MA: Harvard University Press; 1991. [ Google Scholar ]
  • Snowling M, Hulme C. The nature and classification of reading disorders: a commentary on proposals for DSM-5. Journal of Child Psychology and Psychiatry. 2011 pre-published on-line. [ PMC free article ] [ PubMed ] [ Google Scholar ]
  • Stanovich KE. Toward an interactive-compensatory model of individual differences in the development of reading fluency. Reading Research Quarterly. 1980; 16 :32–71. [ Google Scholar ]
  • Terry NP. Dialect variation and phonological knowledge: Phonological representations and metalinguistic awareness among beginning readers who speak nonmainstream American English. Applied Psycholinguistics in press. [ Google Scholar ]
  • Terry NP, Scarborough HS. The phonological hypothesis as a valuable framework for studying the relation of dialect variation to early reading skills. In: Brady S, Braze D, Fowler C, editors. Explaining Individual Differences in Reading: Theory and Evidence. New York, NY: Taylor & Francis Group; 2012. [ Google Scholar ]
  • Thompson CA. Unpublished doctoral dissertation. University of Michigan; 2003. The oral vocabulary abilities of skilled and unskilled African American readers. [ Google Scholar ]
  • Thompson CA, Craig HK, Washington JA. Variable production of African American English across oracy and literacy contexts. Language, Speech, and Hearing Services in Schools. 2004; 35 :269. [ PubMed ] [ Google Scholar ]
  • Tobias S, Duffy TM, editors. Constructivist Theory Applied to Instruction: Success or Failure? New York: Routledge, Taylor and Francis; 2009. [ Google Scholar ]
  • Trudgill P, Chambers JK. Dialects of English: Studies in grammatical variation. London & New York: Longman; 1991. [ Google Scholar ]
  • Vanneman A, Hamilton L, Anderson JB, Rahman T. Statistical Analysis Report NCES 2009-455. National Center for Education Statistics; 2009. Achievement Gaps: How Black and White Students in Public Schools Perform in Mathematics and Reading on the National Assessment of Educational Progress. [ Google Scholar ]
  • Washington J. Issues in assessing the language abilities of African-American children. In: Kahmi A, Pollack K, Harris J, editors. Communication development and disorders in African American children. Baltimore, MD: Paul H. Brookes Publishing Co; 1996. [ Google Scholar ]
  • Willis J. Engaging the whole child: The neuroscience of joyful education. educational leadership. 2007 On-line only: http://www.ascd.org/publications/educational-leadership/summer07/vol64/num09/The-Neuroscience-of-Joyful-Education.aspx .
  • Wolfram W, Schilling-Estes N. American English: dialects and variation. Malden, MA: Blackwell; 1998. [ Google Scholar ]
  • Share full article

Advertisement

Supported by

‘The More You Know About Something the Less Scary It Is’: The Week 7 Winner of Our Summer Reading Contest

Michelle Zhibing Zhou, 15, writes about a Science article highlighting new research that “sent ripples throughout the leech research community and shivers down my spine.”

reading research articles

By The Learning Network

For 15 years, our Summer Reading Contest has been inviting teenagers around the world to tell us what New York Times pieces get their attention and why. This year, for the first time, students can submit either written comments or 90-second video responses.

In the seventh week of our 10-week challenge, we received 1,082 entries, and we list the finalists below. Scroll down to read the work of our winner, Michelle Zhibing Zhou , and to take a look at the variety of topics that caught these students’ eyes, including a nine-year-old chess prodigy, Kamala Harris’s cooking, Alzheimer’s disease, jewelry, cows, anchovies, and preserving memories of the Holocaust.

You can read the work of all of our winners since 2017 in this column , and you can participate in the contest any or every week this summer until Aug. 16. Just check the top of this page , where we post updates, to find the right place to submit your response.

Michelle Zhibing Zhou, 15, from Hong Kong, read the Science article, “ Videos Show That Leeches Can Jump in Pursuit of Blood .” Here is her response:

Leeches! Bleh! Famed for their role in medieval medicine, they’re often dismissed as repulsive, blood-thirsty parasites. So, imagine my surprise and morbid curiosity when I stumbled upon an article that added insight into their world and abated some of my irrational fears. Surprisingly, leeches reminded me of something very different: myself. When I have a craving, like for chocolate, I scurry to the store. Similarly, when motivated by their craving for blood, leeches are “provoked [into] acts of startling athleticism,” springing through the air, then splatting on the ground comically. Adding even more intrigue, the video included ground-breaking footage ending the age-old debate about whether leech-leaping was even possible. The finding sent ripples throughout the leech research community and shivers down my spine. As with many things in life, the more you know about something the less scary it is. The article helped dissipate some of my fear and disgust, presenting leeches as being motivated by food (like me!), and with amusing behaviors. The article changed my perception of leeches from nightmare fuel to riveting little creatures. Don’t get me wrong though, while leeches are endlessly captivating, being that, to them, I’m a tasty walking bag of blood, I don’t want them snacking on me in a tropical forest. I’ll stick to watching them from afar, behind a screen, and preferably while nibbling on a bar of chocolate.

In alphabetical order by the writer’s first name.

Nazira Musabaeva on “ After 12 Years of Reviewing Restaurants, I’m Leaving the Table ”

Noa Riss on “‘ Crown Jewels of the Jewish People’: Preserving Memories of the Holocaust ”

Shi Yi Yang on “ A Family Dinner with My Wife and Girlfriend ”

Sissi Ma on “ When It Comes to Food and Politics, Kamala Harris Is Riffing on the Recipe ”

Honorable Mentions

Alex Ding on “ The Youngest Pandemic Children Are Now in School, and Struggling ”

We are having trouble retrieving the article content.

Please enable JavaScript in your browser settings.

Thank you for your patience while we verify access. If you are in Reader mode please exit and  log into  your Times account, or  subscribe  for all of The Times.

Thank you for your patience while we verify access.

Already a subscriber?  Log in .

Want all of The Times?  Subscribe .

  • Future Students
  • Current Students
  • Faculty/Staff

Stanford GSE

News and Media

  • News & Media Home
  • Research Stories
  • School’s In
  • In the Media

You are here

New study explores what makes digital learning products more – or less – effective.

Group of early elementary age students working on tablets

Educational technology has become a fixture in the U.S. classroom, but scholars continue to debate its effectiveness – some even arguing that the products might deter learning by taking students’ time and attention away from more powerful supports. 

What does research show about the effectiveness of edtech? Does the impact vary when it comes to teaching certain skills and student populations? How can schools determine which products are most useful for their own setting and purposes? 

A new Stanford-led study sheds light on the value of edtech interventions, with a focus on products aimed at helping elementary school students develop early reading skills. In a meta-analysis of studies conducted over the past two decades, the researchers found that the effectiveness of tech products varied considerably, depending on particular features of the interventions and the skills they targeted. 

“When we talk about digital learning products, they’re really not all the same – there’s a wide range,” said Rebecca Silverman , the Judy Koch Professor of Education at Stanford Graduate School of Education (GSE), a faculty affiliate of the Stanford Accelerator for Learning , and the study’s lead author. “There isn’t a single answer to whether digital technologies support literacy. The question is much more complex: Which products, with which characteristics, under which conditions?” 

The paper , published July 31 in the peer-reviewed journal Review of Educational Research , was co-authored by Elena Darling-Hammond, a doctoral student at the GSE; Kristin Keane, a postdoctoral scholar at the GSE; and Saurabh Khanna, PhD ’23, who is now an assistant professor at the University of Amsterdam. 

Rebecca Silverman

Stanford GSE Professor Rebecca Silverman

Accounting for variability

For the meta-analysis, the researchers drew on 119 studies published between 2010 and 2023 to examine the use of various digital interventions in kindergarten through fifth grade, including computer programs, e-books, online games, and videos. 

The study is unique, they said, in its focus on edtech at the elementary school level and its review of interventions across four skills: decoding (the ability to read words quickly and accurately), language comprehension (understanding the meaning of words), reading comprehension (processing the meaning of a passage), and writing proficiency (the ability to convey ideas in writing).

Their analysis found positive effects on elementary school students’ reading skills overall, indicating that generally, investing in educational technology to support literacy is warranted. But when the researchers isolated particular learning outcomes to measure effectiveness, they found wide variability, suggesting that the effectiveness of a particular edtech product can depend on different factors, including features of the tool and characteristics of the users.

The authors observed that most studies – and the majority of products in the marketplace – focused on basic decoding, where students use phonetic skills to understand the relationship between written letters and their sounds. Relatively few studies considered language and reading comprehension, and only a handful looked at writing proficiency. 

“Decoding is a fairly constrained construct involving a relatively circumscribed set of skills,” Silverman said. “There are only so many letters and sounds and letter-sound combinations that kids need to learn, so it’s generally easier to teach and see change over time.”

Language comprehension is a more complex construct, she said, involving a vast number of concepts, word meanings, and sentence constructions and the ability to make connections and build knowledge. “Its complexity makes it harder to teach and see progress. But it’s a crucial skill to be able to access texts and content, so we need more tools and research focused on that piece.” 

Product features that appeared to account for some of the variability in effectiveness included the type of technology, the duration of the intervention, and the instructional approach (that is, whether it emphasized repetition and facts, strategies to organize and process information, or open-ended tasks). 

The analysis found, for example, that certain personalization, gamification, and interactive feedback features, like pop-up questions and clickable definitions, were not effective for supporting more complex skills like reading comprehension.  

Where student characteristics were concerned, socioeconomic status surfaced as one factor moderating effectiveness: With decoding as an outcome, for example, studies with a substantial percentage of students from low socioeconomic backgrounds tended to have larger effects compared with other studies, which Silverman said could be due to the programs they used being more geared toward their needs. 

The researchers suspected that disability and language status would also emerge as a factor in the variability they uncovered, but few studies disaggregated findings based on these backgrounds. 

“A program might not benefit some kids as much as others, and if we don’t track that in a systematic way, we’re not going to know,” Silverman said. “Right now, it’s not being systematically captured in the research, and that’s a problem.” 

The researchers also noted that few studies addressed edtech’s impact on students’ motivation or engagement, and few included follow-up over time, to assess whether the effects lasted months or even years after the intervention. 

Considerations for school leaders

The findings point to several directions for educators and policymakers, the researchers concluded. For one thing, Silverman said, districts contemplating a particular product should carefully consider whether it’s appropriate for their population of students, and whether the content and approach aligns with the curriculum and classroom teaching. 

She advised that, rather than taking marketing claims at face value, districts conduct a critical analysis of any program before deciding whether to adopt it for their schools. “Is it following the principles of effective practice for the skills you’re targeting with that program?” she said. “What studies have been done on it? How strong is the company’s own research? Has anybody done any independent research?”

Districts can also generate their own data, for example, by running a pilot program in which some schools or classrooms implement an edtech intervention, comparing their outcomes against the schools that don’t. “You may not be able to isolate [the effects of the program] completely,” Silverman said, “but an analysis can suggest whether this product is helpful.”

If a product doesn't appear to produce positive effects, districts can partner with researchers to try to figure out why — or they can move on to trying other tools and evaluate those, she said. “We don’t want kids to keep using products that aren’t helpful.”  

More Stories

Children with teacher in an elementary school classroom

⟵ Go to all Research Stories

Get the Educator

Subscribe to our monthly newsletter.

Stanford Graduate School of Education

482 Galvez Mall Stanford, CA 94305-3096 Tel: (650) 723-2109

  • Contact Admissions
  • GSE Leadership
  • Site Feedback
  • Web Accessibility
  • Career Resources
  • Faculty Open Positions
  • Explore Courses
  • Academic Calendar
  • Office of the Registrar
  • Cubberley Library
  • StanfordWho
  • StanfordYou

Improving lives through learning

reading research articles

  • Stanford Home
  • Maps & Directions
  • Search Stanford
  • Emergency Info
  • Terms of Use
  • Non-Discrimination
  • Accessibility

© Stanford University , Stanford , California 94305 .

All Articles Education Insights The science of reading: A formula to great literary instruction

The science of reading: A formula to great literary instruction

Explicit reading instruction using the science of reading can make an enormous difference in students' ability to read, Kymyona Burk writes.

By Kymyona Burk 08/07/24

Education Insights

Adorable Hispanic schoolgirl smiles while delight as she reads a picture book while standing in the school library. for article on science of reading

(SDI Productions/Getty Image)

For most Americans, learning to read is a fundamental part of early childhood. Whether it was our parents reading to us at night or early school lessons full of colorful and rhyming stories, reading is a part of the beautiful and complicated puzzle that makes us who we are. 

Our brains have been trained for years to take in the information that we see and turn it into something we understand. For students, the most crucial period for this training takes place between kindergarten and third grade. During those years, students slowly transition from learning to read to reading to learn. By fourth grade, most curriculum is being taught through reading. 

The way we teach early literacy plays a huge role in a student’s ability to successfully read by the crucial fourth-grade checkpoint. 

Flawed instruction can lead to reading failures

America is facing a reading crisis. According to NAEP , only 33% of fourth-grade students in the U.S. are reaching grade-level proficiency in reading.

This number is so low partly because schools have been using harmful reading practices such as three-cueing. This flawed instruction model teaches students to read based on meaning, structure and visual cues, redirecting the student from the word itself. Essentially, students are taught to guess a word based on a picture instead of sounding out the word. 

Eight states banned three-cueing in 2023. 

So, what is the right way to teach students to read? ExcelinEd is one of many education-focused entities that embrace and promote the research and evidence-based practice known as the science of reading.

To ensure every child can read by third grade, all educators must be trained in the science of reading, an evidence-based approach that teaches phonemic awareness, phonics/decoding, fluency, vocabulary, and comprehension. This is the opposite of the three-cueing system. 

Science of reading is the best way forward

The science of reading is a culmination of research that incorporates information learned through studies of developmental psychology, educational psychology, cognitive science and cognitive neuroscience.

For example, neuroscientists using functional MRI scans showed the section of the brain dealing with language was significantly more active in young students learning to read through a phonetic-based curriculum than in kids who were taught by three-cueing methods. 

While the science of reading is based on phonetic learning, other skills such as oral language, alphabet knowledge, phonemic awareness and vocabulary development are critical to becoming a skilled reader.  

Researchers have been able to determine the brain’s path to reading comprehension and the type of explicit instruction that is needed for students who may have difficulty learning to read.

This style of reading instruction is working for students across the country. A  2023 study on select California schools found improvements in reading proficiency using evidence-based approaches.  California doesn’t mandate science of reading instruction, so the study focused on 76 of the state’s lowest-performing schools. Researchers found the science of reading approach raised test scores in English and math for third-graders in 2022 and 2023. The increase in comprehension was the equivalent of attending an additional quarter of a school year compared to other students who did not use evidence-based instruction. 

As of April 2024, 38 states and the District of Columbia have passed laws or implemented new policies related to evidence-based reading instruction. For the majority of those states, that includes training teachers in how to better implement science of reading instruction.

For example, California has adopted 10 of 18 recommended early literacy policy fundamental principles, one of 12 states operating without requiring teachers to learn how to teach using the science of reading.  

Science of reading can make a difference

I’ve seen the change that teaching with explicit instruction and scientifically based methods can make in a state. 

From 2013 to 2019, I was the state literacy director for the Mississippi Department of Education. In 2013, the state ranked last in the nation in fourth-grade reading, according to NAEP. We set out to change that.

With courageous leadership from then-state Superintendent Carey Wright, Ed.D., (now state superintendent for Maryland) and policy know-how from state officials, we were able to implement the Literacy-Based Promotion Act to mandate research-based literary instruction and better professional learning opportunities for Mississippi teachers. 

As of 2023, Mississippi ranks 21st   in the nation for fourth-grade reading. Some call this change the Mississippi Miracle . It’s not a miracle . It’s proof that scientifically research-based instruction works.   

Evidence-based literary instruction is the key to solving our nation’s reading crisis. Working together, as other states have done over the past two decades, we can bring more science of reading methods to our classrooms to ensure our youngest readers have a strong foundation to build on as they move through and beyond school.   

Opinions expressed by SmartBrief contributors are their own. 

_________________________

Subscribe to SmartBrief’s FREE email ASCD newsletter to see the latest hot topics in education. It’s among SmartBrief’s more than 250 industry-focused newsletters .

IMAGES

  1. Reading a Scholarly Article

    reading research articles

  2. Reading Scholarly Articles

    reading research articles

  3. How to Read a Journal Article in 7 Steps (2024)

    reading research articles

  4. How to Read a Research Paper

    reading research articles

  5. Research Survival Guide

    reading research articles

  6. A Beginner's Guide to Reading Research Articles

    reading research articles

COMMENTS

  1. Art of reading a journal article: Methodically and effectively

    Reading scientific literature is mandatory for researchers and clinicians. With an overflow of medical and dental journals, it is essential to develop a method to choose and read the right articles.To outline a logical and orderly approach to reading ...

  2. How to read and understand a scientific paper

    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.

  3. How to Read a Scholarly Article

    Identify the different parts of a scholarly article. Efficiently analyze and evaluate scholarly articles for usefulness. This page will focus on reading scholarly articles — published reports on original research in the social sciences, humanities, and STEM fields. Reading and understanding this type of article can be challenging.

  4. ENGL105

    How to Read a Scholarly Article. Scholarly articles can be very intimidating! They are written by experts for other experts in a field, so they can be filled with technical, confusing jargon.

  5. Ten simple rules for reading a scientific paper

    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 ...

  6. <em>Reading Research Quarterly</em>

    The simple view of reading is commonly presented to educators in professional development about the science of reading. The simple view is a useful tool for conveying the undeniable importance—in fac...

  7. Reading Comprehension Research: Implications for Practice and Policy

    Learn how reading comprehension research can inform educational practice and policy, and explore the latest trends and challenges in the field.

  8. How to (seriously) read a scientific paper

    Adam Ruben's tongue-in-cheek column about the common difficulties and frustrations of reading a scientific paper broadly resonated among Science Careers readers. Many of you have come to us asking for more (and more serious) advice on how to make sense of the scientific literature, so we've asked a dozen scientists at different career stages and in a broad range of fields to tell us how they ...

  9. Infographic: How to read a scientific paper

    Much of a scientist's work involves reading research papers, whether it's to stay up to date in their field, advance their scientific understanding, review manuscripts, or gather information for a project proposal or grant application. Because scientific articles are different from other texts, like novels or newspaper stories, they should be read differently.

  10. Reading Research Effectively

    Offers detailed guidance on how to develop, organize, and write a college-level research paper in the social and behavioral sciences.

  11. PDF How to Read a Paper

    Researchers spend a great deal of time reading research pa-pers. However, this skill is rarely taught, leading to much wasted e ort. This article outlines a practical and e cient three-pass method for reading research papers. I also de-scribe how to use this method to do a literature survey.

  12. Reading a Scientific Article

    Reading a Scientific Article Attempting to read a scientific or scholarly research article for the first time may seem overwhelming and confusing. This guide details how to read a scientific article step-by-step. First, you should not approach a scientific article like a textbook— reading from beginning to end of the chapter or book without pause for reflection or criticism. Additionally, it ...

  13. How to Read Scholarly Articles: Strategies for Reading

    Here you will learn how to recognize a scholarly article, its components and strategies for reading

  14. Library Research Guides: STEM: How To Read A Scientific Paper

    Reading a Scientific Paper Reading a scientific paper can seem like a daunting task. However, learning how to properly read a scholarly article can make the process much easier! Understanding the different parts of a scientific article can help the reader to understand the material.

  15. Journal of Research in Reading

    Published on behalf of the United Kingdom Literacy Association, Journal of Research in Reading is principally devoted to reports of original empirical research in reading and closely related fields (e.g., spoken language, writing), and to informed reviews of relevant literature. We publish papers from researchers in any relevant field on the ...

  16. Reading Research Effectively

    Reading a Scholarly Article or Research Paper Reading Research Publications Effectively It's easy to feel overwhelmed and frustrated when first reading a scholarly article or research paper. The text is dense and complex and often includes abstract or convoluted language.

  17. How the Science of Reading Informs 21st‐Century Education

    The science of reading should be informed by an evolving evidence base built upon the scientific method. Decades of basic research and randomized controlled trials of interventions and instructional routines have formed a substantial evidence base to guide best practices in reading instruction, reading intervention, and the early identification ...

  18. Ten simple rules for reading a scientific paper

    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 ...

  19. Reading Research Quarterly

    Reading Research Quarterly. An official journal of the International Literacy Association, Reading Research Quarterly is the leading global journal offering multidisciplinary scholarship on literacy among learners of all ages, including the latest research studies (methods, results, effects, findings, and implications).

  20. <em>Reading Research Quarterly</em>

    When preparations began for this special issue of Reading Research Quarterly, the term science of reading (SOR) was a dominant part of discourse in education. We kept hearing the term in discussions across our experiences as researchers, teacher educators, partners of local education agencies, community members, and even parents. Even the media weighed in. Those invested saw the SOR as either ...

  21. The Science of Reading and Its Educational Implications

    Abstract. Research in cognitive science and neuroscience has made enormous progress toward understanding skilled reading, the acquisition of reading skill, the brain bases of reading, the causes of developmental reading impairments and how such impairments can be treated. My question is: if the science is so good, why do so many people read so ...

  22. Critical Issues in the Science of Reading: Striving for a Wide-Angle

    This report reflects a panel presentation and discussion at the 2020 Literacy Research Conference focused on the science of reading (SoR). Each panelist present...

  23. 'The More You Know About Something the Less ...

    Michelle Zhibing Zhou, 15, writes about a Science article highlighting new research that "sent ripples throughout the leech research community and shivers down my spine."

  24. The Science of Reading Comprehension Instruction

    Abstract Decades of research offer important understandings about the nature of comprehension and its development. Drawing on both classic and contemporary research, in this article, we identify some key understandings about reading comprehension processes and instruction, including these: Comprehension instruction should begin early, teaching word-reading and bridging skills (including ...

  25. New study explores what makes digital learning products more

    Research led by Stanford Professor Rebecca Silverman analyzes studies on edtech interventions for early reading skills.

  26. Reading Research Quarterly: Volume 55, Issue S1

    How the Reading for Understanding Initiative's Research Complicates the Simple View of Reading Invoked in the Science of Reading Gina N. Cervetti, P. David Pearson, Annemarie S. Palincsar, Peter Afflerbach, Panayiota Kendeou, Gina Biancarosa, Jennifer Higgs, Miranda S. Fitzgerald, Amy I. Berman Pages: S161-S172 First Published: 30 August 2020

  27. The science of reading: A formula to great literary instruction

    Explicit reading instruction using the science of reading can make an enormous difference in students' ability to read, Kymyona Burk writes.